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
|
---|---|---|---|---|---|---|---|---|---|
1711.06986 | 1 | 1711 | 2017-11-19T09:00:16 | Numerical-experimental observation of shape bistability of red blood cells flowing in a microchannel | [
"physics.bio-ph",
"physics.comp-ph",
"physics.flu-dyn"
] | Red blood cells flowing through capillaries assume a wide variety of different shapes owing to their high deformability. Predicting the realized shapes is a complex field as they are determined by the intricate interplay between the flow conditions and the membrane mechanics. In this work we construct the shape phase diagram of a single red blood cell with a physiological viscosity ratio flowing in a microchannel. We use both experimental in-vitro measurements as well as 3D numerical simulations to complement the respective other one. Numerically, we have easy control over the initial starting configuration and natural access to the full 3D shape. With this information we obtain the phase diagram as a function of initial position, starting shape and cell velocity. Experimentally, we measure the occurrence frequency of the different shapes as a function of the cell velocity to construct the experimental diagram which is in good agreement with the numerical observations. Two different major shapes are found, namely croissants and slippers. Notably, both shapes show coexistence at low (<1 mm/s) and high velocities (>3 mm/s) while in-between only croissants are stable. This pronounced bistability indicates that RBC shapes are not only determined by system parameters such as flow velocity or channel size, but also strongly depend on the initial conditions. | physics.bio-ph | physics | Numerical-experimental observation of shape bistability of red blood cells flowing in a
microchannel
Achim Guckenberger,1 Alexander Kihm,2 Thomas John,2 Christian Wagner,2, 3 and Stephan Gekle1
1Biofluid Simulation and Modeling, Fachbereich Physik, Universität Bayreuth, Bayreuth
2Experimental Physics, Saarland University, 66123, Saarbrücken, Germany
3Physics and Materials Science Research Unit, University of Luxembourg, Luxembourg, Luxembourg
(Dated: November 19, 2017)
7
1
0
2
v
o
N
9
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
6
8
9
6
0
.
1
1
7
1
:
v
i
X
r
a
Red blood cells flowing through capillaries assume a wide variety of different shapes owing to
their high deformability. Predicting the realized shapes is a complex field as they are determined
by the intricate interplay between the flow conditions and the membrane mechanics. In this work
we construct the shape phase diagram of a single red blood cell with a physiological viscosity ratio
flowing in a microchannel. We use both experimental in-vitro measurements as well as 3D numerical
simulations to complement the respective other one. Numerically, we have easy control over the
initial starting configuration and natural access to the full 3D shape. With this information we obtain
the phase diagram as a function of initial position, starting shape and cell velocity. Experimentally,
we measure the occurrence frequency of the different shapes as a function of the cell velocity to
construct the experimental diagram which is in good agreement with the numerical observations.
Two different major shapes are found, namely croissants and slippers. Notably, both shapes show
coexistence at low (< 1 mm/s) and high velocities (> 3 mm/s) while in-between only croissants are
stable. This pronounced bistability indicates that RBC shapes are not only determined by system
parameters such as flow velocity or channel size, but also strongly depend on the initial conditions.
I.
INTRODUCTION
Red blood cells (RBCs) are the major constituent of
mammalian blood and therefore determine the majority
of its flow properties. One of the most amazing features
of RBCs is their deformability, allowing them to squeeze
through channels with diameters much smaller than their
own equilibrium size [1–3]. Another consequence of their
deformability is the wide range of stationary and non-
stationary shapes assumed by the RBCs in microchannel
flows with dimensions similar to or slightly larger than
the RBC equilibrium radius [4–6]. Understanding and
being able to predict these shapes is of high importance
for a variety of reasons. From a fundamental point of
view, it serves as the foundation in a bottom-up approach
to understand the properties of red blood cell suspensions
which are chiefly determined by single particle behavior
[7–12]. From an applied perspective, a series of recent
investigations have devised promising approaches for sort-
ing cells based on their mechanical properties either in
lateral displacement devices [13] or using high-speed video
microscopy [14]. Finally, knowledge of the precise cell
shape is also essential for accurately measuring geometric
properties of cells [15].
The most frequently observed shapes of RBCs in mi-
crochannel flows are the so-called "croissant" and "slipper"
shapes. Examples are depicted in figure 1. Some re-
searchers refer to croissants also as parachutes, although
here we prefer the term croissant since our shapes are
not perfectly rotationally symmetric (similar to the ones
found by Farutin and Misbah [16]). Probably one of the
earliest experimental study on isolated red blood cells
in flow was performed by Gaehtgens et al. [17], where
slippers as well as parachutes have been found depend-
ing on the diameter of the cylindrical channel. Suzuki
et al. [18] presented a phase diagram of parachutes and
slippers as a function of velocity and confinement in a
cylindrical tube. Slippers dominated at smaller diameters
and higher velocities. Secomb et al. [19] compared experi-
ments with 2D simulations in cylindrical channels of 8 µm
diameter for a cell velocity of approximately 1.25 mm/s.
Furthermore, two other publications [20, 21] considered
the flow of RBCs at very low viscosity ratios of λ (cid:46) 0.27.
They presented a phase diagram showing parachutes and
slippers, where the velocity was varied in the very high
regime of 10 to 170 mm/s. Tomaiuolo et al. [22] found
parachutes at smaller and slippers at higher velocities
in cylindrical channels of 10 µm diameter. A subsequent
study [23] as well as Prado et al. [24] considered the tran-
sient during start-up of the flow. Cluitmans et al. [25]
detected croissants at lower ((cid:46) 5 mm/s) and slippers at
higher velocities ((cid:38) 10 mm/s) in rectangular channels
with widths (cid:54) 10 µm. Moreover, Quint et al. [26] found a
stable slipper and a metastable croissant at the same set
of parameters in a wider channel of 25 µm× 10 µm. Other
publications presenting experiments in channel flow also
touch the subject of RBC shapes but focus on other as-
pects such as the methodology [27–34], dense suspensions
and cell interactions [17, 21, 34–41] or use vastly larger
channel diameters [12, 42].
Numerical simulations and semi-analytical calculations
of isolated particles in microchannels mostly studied ax-
isymmetric RBCs [43–45] or 2D vesicles [5, 6, 46–51]. The
numerical work by Aouane et al. [5], for example, identi-
fied a large amount of dynamics including deterministic
chaos. The first full 3D simulation of single cells with a
realistic RBC model (but with a ratio of inner to outer
viscosity of λ = 1) was conducted by Noguchi and Gomp-
per [52] who used a cylindrical tube with a diameter of
9.2 µm. They found the typical discocyte shape below
(a)
(b)
(c)
(d)
(e)
(f)
2
(g)
Figure 1. Typical RBC shapes from simulations and experiments. (a) The typical discocyte shape employed in some of the
simulations as the starting shape. Half of it was made transparent for illustration purposes. Its horizontal diameter is 8 µm.
(b) A typical croissant observed in the experiments when applying a pressure drop of 100 mbar (cell velocity (0.98 ± 0.07) mm/s).
(c) A croissant with a velocity of ≈ 1.1 mm/s obtained from the numerical simulations. (d) The cross-section of the croissant
from (c). (e) A slipper from the experiments at 500 mbar (cell velocity (5.16 ± 0.11) mm/s). (f) A typical slipper from the
simulations with a cell velocity of ≈ 5.2 mm/s. (g) The cross-section of the slipper from (f). The black lines on the shapes from
the simulations depict the mesh. The bottom and top black lines in all figures are the walls (Ly ≈ 12 µm apart), while the small
black lines are scale bars of length 2 µm. The flow is in the positive x-direction (except in figure (a) where no flow exists).
and parachutes above a critical velocity which depends
on the elastic parameters. A subsequent study by the
same group additionally explored this threshold as a func-
tion of confinement [53]. Moreover, Fedosov et al. [4]
presented very detailed phase-diagrams where the veloc-
ity and confinement was varied for three different sets
of elastic moduli and a viscosity ratio of λ = 1. They
observed four distinct regions where snaking, tumbling,
slippers and parachutes occurred. Recently, Ye et al. [54]
considered the shapes of an RBC with λ = 1 in rectan-
gular microchannels (with width 10 µm and aspect ratios
1 to 2) for the three cell velocities 4, 20 and 100 mm/s
and observation times up to ≈ 0.03 s. Snapshots after
this short initial transient showed parachutes or slightly
slipper-like shapes.
Bistability, i.e. the observation of two different stable
shapes depending on the initial condition but at otherwise
identical system parameters, was barely considered so far.
It was observed only numerically for simpler situations
such as close-to-spherical vesicles in unbounded Poiseuille
flow [16] or near a single wall [55], for a 2D RBC model
in bounded Poiseuille flow [19], for the initial transient
of a red blood cell in a rectangular channel [54] or for
simple shear flows [12, 56–58]. No systematic experimen-
tal investigations exist for cells flowing in microchannels.
Moreover, the 3D simulations and experimental investiga-
tions that were mentioned above and that consider the
RBC shapes in microchannels in more detail all used a
viscosity ratio of λ (cid:54) 1, although 2D simulations showed
that choosing a physiologically more realistic value of
λ ≈ 5 [59] can significantly affect RBC dynamics [6, 49].
Here we present a detailed systematic experimental-
numerical study on the steady-state shape of isolated
red blood cells in a rectangular microchannel. We use
the physiological viscosity ratio of λ = 5 appropriate for
healthy human red blood cells in the microcirculation [59].
The initial position is varied in the simulations directly,
while experimentally we determine it via measurements
at the channel entrance. Our central finding is that the
initial starting position of the RBC has a decisive influence
on the final steady-state shape of the red blood cell.
We begin by outlining our experimental and numerical
methods in section II. Afterwards, the results from our
experiments (section III) and simulations (section IV) are
presented, while section V is dedicated to their detailed
comparison. Finally, we conclude our work in section VI.
II. METHODS
A. Experimental setup
The sample preparation and experimental setup is
mostly identical to the one used recently by Clavería
et al. [35]. In short, human red blood cells were obtained
from healthy donors by needle-prick and used within
three hours. After appropriate preparation [35], they
are suspended in a phosphate buffered saline (PBS) and
bovine serum albumin solution which has a viscosity of
approximately 1 mPa s. The viscosity ratio of the cells is
therefore λ ≈ 5 [26]. This value corresponds to the typi-
cal physiological value of healthy red blood cells in blood
plasma [59]. The RBCs are pumped through rectangular,
PDMS-based channels by a high-precision pressure device
(Elveflow OB 1, MK II) with pressure drops ranging from
20 to 1000 mbar at room temperature. The channels have
a cross-section width of Ly = (11.9 ± 0.3) µm and a height
of Lz = (9.7 ± 0.3) µm without any applied pressure drop
and are thus similar to the vessel diameters found in the
microvascular system [60]. We use rectangular rather than
cylindrical channels since they are easier to manufacture,
are therefore prevalent in lab-on-a-chip devices and have
the merit that cells are not rotated randomly around their
axis due to the missing rotational symmetry. The latter
property greatly simplifies the microscopic observation
and analysis of the RBCs.
The hematocrit (volume percentage of RBCs) in the
reservoir before the inlet is always (cid:46) 1.0 %, i.e. very low.
Nevertheless, we find cells flowing in clusters as well as
single cells. For the present work we have analyzed only
the latter. To this end, previous experimental and theo-
retical results showed that the hydrodynamic interaction
yxin a linear channel decays exponentially, and becomes
negligible if the inter-particle distance is more than twice
the channel width [53, 61, 62]. Considering that our chan-
nel has the dimensions ≈ 12 µm × 10 µm, cells can be
considered as being single for distances (cid:38) 25 µm. We
only used cells that were at least 40 µm apart from other
entities.
We perform measurements at two locations along the
channel, namely at the entrance (x = 0 mm) and at
x = 10 mm downstream. Vessel lengths in-between bifur-
cations in the microvascular system are less than 1 mm,
i.e. much shorter [63]. Nevertheless, this is not necessar-
ily true for in-vitro experiments or lab-on-a-chip devices,
and the long-time behavior also holds information about
the general intrinsic properties. The flowing RBCs are
recorded by an inverted bright-field microscope (Nikon
TE 2000-S) with an oil-immersion objective (Nikon CFI
Plan Fluor 60×, NA = 1.25) and a high-resolution camera
(Fastec HiSpec 2G) at a frame rate of 400 frames per sec-
ond. The camera is aligned along the z-direction so that
the photographs show the cells in the x-y-plane (compare
figure 2). Hence, determination of the z-position is not
possible, but also not absolutely necessary as our simula-
tions always show a z-position of nearly 0 (see section V).
We analyze the recorded image sequence with a custom
MATLAB script that detects each projected cell shape
and the corresponding 2D center of mass position. It ad-
ditionally tracks the cell position over the image sequence
to obtain the individual cell velocity. Considering the
optical setup, we assume an uncertainty in the position
measurements of ±sP with sP = 0.1 µm.
B. Simulation setup
The numerical simulations mimic our experimental
setup as far as possible. Hence, we place a single red
blood cell in a rectangular channel as shown in figure 2.
The channel has a cross-section of width Ly = 12 µm
and height Lz = 10 µm. Periodic boundary conditions
are assumed in the x-direction with a periodicity of
Lx = 42.7 µm, in agreement with above estimates for
the decay of hydrodynamic interactions.
simply given by rinit = (cid:112)y2
We vary the initial y-z-position (relative to the chan-
nel center) of the RBC's centroid along the line zinit =
5yinit/9, which almost corresponds to the channel diag-
onal. The corresponding initial radial position is thus
. When starting
with the typical discocyte equilibrium shape [64, 65], as
depicted in figure 1(a), the RBC axis is aligned with the
channel axis (as shown in figure 2). Cell velocities are
extracted by considering the difference of the centroids
between successive time steps. During the simulation,
we monitor several quantities such as the radial, y- and
z-positions, the RBC asphericity or the cell velocity as
well as the full 3D shape to determine when a steady state
has been reached.
init + z2
init
Regarding the actual modeling of the constituents, the
3
Figure 2. Simulation setup: A single red blood cell is placed
in a rectangular channel of width Ly = 12 µm and height
Lz = 10 µm. Periodic boundary conditions are employed.
Initially, the centroid of the cell is offset from the center axis
along the left black arrow by a distance rinit. The depicted
RBC illustrates the discocyte starting shape, although other
shapes have been used, too. Furthermore, the black lines on
the surfaces illustrate the employed meshes. The arrow at the
top shows the view from the camera in the experiments (i.e.
onto the x-y-plane) and the flow is in the positive x-direction.
RBC is filled with a Newtonian fluid with a dynamic
viscosity µRBC, whereas the ambient flow is a Newtonian
fluid with the dynamic viscosity µ = 1.2 × 10−3 kg/(s m)
of blood plasma [66–68]. We set the viscosity ratio λ =
µRBC/µ to a value of 5 in all simulations. The surface
area of the RBC is set to 140 µm2 and the volume is
set to 100 µm3 (see e.g. references 67 and 69), leading to
a large radius of RRBC = 4 µm when the cell is in the
typical discocyte equilibrium shape (figure 1(a)). The
mechanics of the infinitely thin membrane are governed
by Skalak's law [70, 71] for the in-plane elasticity with
a shear modulus of κS = 5 × 10−6 N/m [72, 73] and an
area dilatation modulus of κA = 100 κS. This value for
κA ensures that the area changes remain below 2 % in all
cases. We take the reference state for the Skalak model to
be the typical discocyte shape [64, 65]. The membrane is
additionally endowed with some bending resistance which
is modeled according to the Canham-Helfrich law [74–76],
where the bending modulus is fixed to κB = 3 × 10−19 N m
[73, 77]. The spontaneous curvature is set to zero.
We use 2048 flat triangles to discretize the RBC in our
numerical implementation. The forces are computed as
described by Guckenberger et al. [78], with Method C
therein being used for the bending contribution. An un-
avoidable artificial volume drift of the cell is countered by
adjusting the velocity to obey the no-flux condition and
by a subsequent rescaling of the object [79, 80]. Moreover,
the channel is represented by 2166 flat triangles. The
corners are rounded to prevent numerical problems (com-
pare figure 2). Rather than prescribing a zero velocity at
the channel walls, we use a penalty method for efficiency
reasons with a spring constant of κW = 1.9 × 107 N/m3
[6, 80]. Increasing the triangle counts and the box length
Lx did not change the results significantly.
The Reynolds number in the considered system is
defined as Re = 2RRBCumaxρ/µ. For a velocity of
umax (cid:54) 10 mm/s and the density ρ ≈ 103 kg/m3 of the am-
bient and inner liquid we therefore have Re < 0.1. Hence,
the flow can be appropriately described using the Stokes
equation. This allows us to employ the boundary integral
method (BIM) [81] for 3D periodic systems [80, 82]. Note
that this method requires to prescribe a certain average
flow through the whole unit cell instead of a pressure
drop within the channel. The latter is unfortunately not
easily accessible. We therefore compare with experiments
by means of cell velocities. Continuing, the integrals
are computed by a standard Gaussian quadrature with
7 points per triangle in conjunction with linear inter-
polation of nodal quantities and appropriate singularity
removal for the single- and double-layer potentials [80].
Furthermore, we use the smooth particle mesh Ewald
(SPME) method [83] to accelerate the computation of the
periodic Green's functions; cutoff errors are kept below
5 × 10−5. The resulting linear system is solved via GM-
RES [84] up to a residuum of 10−5, and the kinematic
condition is integrated in time using the adaptive Bogacki-
Shampine algorithm [85] with the absolute tolerance set
to 10−5RRBC. When the run-times are normalized to a
two-socket system with 28 cores, each simulation took 1
to 29 days, with an average of around 5 days. The phase
diagrams below are formed by 329 of such simulations in
total. Further details on the numerical method as well as
verifications of the implementation can be found in our
previous publications [26, 78, 80, 86].
III. EXPERIMENTAL RESULTS
We classify cells in the experiments either as croissants,
slippers or "other" not uniquely identifiable or completely
different shapes. Typical slipper and croissant shapes are
shown in the photographs (b) and (e) of figure 1. See
the supplementary information (SI) for a collection of all
images.
To systematically investigate the occurrence of the dif-
ferent shapes, we vary the imposed pressure drops from
20 to 1000 mbar. The corresponding cell velocities range
from 0.14 mm/s to 10.6 mm/s, covering the whole physio-
logical range in microchannels [60, 87, 88]. We consider
the cells 10 mm away from the channel entrance where
most of the cells reached a steady state [35]. Figure 3(a)
depicts the fraction of observed shapes as a function of the
measured cell velocities, constituting our central result
from the experiments. This distribution was obtained
by considering typically more than 100 cells per imposed
pressure drop. The average velocities were computed by
averaging over all cells at a certain pressure drop, with
the horizontal error bars showing the corresponding stan-
dard deviations σu in cell velocity. Not all velocities are
the same because croissants and slippers have different
velocities at otherwise identical flow conditions [26], and
4
because of the natural variations of cell properties such
as elasticity and size, as also noted by Tomaiuolo et al.
[22]. See the supplementary information for more de-
tails. Considering figure 3(a), high velocities obviously
favor slippers while croissants are the most prominent
for medium velocities. A pronounced peak exists from
around 1 to 2 mm/s. Very small velocities produce mostly
shapes that fall outside our simple two-state classification.
Figure 3(b) illustrates the corresponding estimated
probability density function of the center of mass y-
position of the cells at the various pressure drops. This
estimate was obtained from the measured y-positions
by using the kernel density estimator as implemented
in MATLAB R2017a (ksdensity) with a support of
[−6, 6] µm and otherwise default settings. Thus, croissants
and "others" occurring at lower velocities are centered in
the channel, while slippers occurring at high velocities
show a pronounced off-centered position. The assumed
shapes therefore imply a certain y-position within the
channel with slippers being off-centered and croissants
centered. This is confirmed when analyzing the offset
distribution separately for each shape class as shown in
the supplementary information.
From figure 3(a) it is tempting to conclude that the flow
velocity is the major parameter that determines the RBC
shape with low velocities favoring centered and high veloc-
ities favoring off-centered flow positions. However, looking
at the cell positions near the channel entrance (figure 4)
we find that already upon entering the channel RBCs are
not homogeneously distributed. At low velocities we ob-
serve a clear bias towards a centered initial position, with
the distribution becoming approximately homogeneous
only at the highest measured velocities. These experi-
mental observations allow two distinct parameters as the
reason for the dominance of the slipper shapes at high
velocities: either the higher flow velocity itself or the more
off-centered entry into the channel. To disentangle these
two possibilities we now present numerical simulations
whose geometry directly corresponds to the experimental
setup.
IV. NUMERICAL RESULTS
We numerically study the behavior of a single RBC in
a rectangular microchannel by varying the imposed flow
velocity, the initial shape and the initial offset rinit from
the centerline of the tube (see section II B). After starting
the flow, we wait until the RBC reaches the steady state
where the shape as well as the radial position does no
longer change, or alternatively until periodic motion is
observed.
In the majority of cases, we observe two different states:
A croissant shape (which moves as a rigid body, figure 1(c))
and a slipper shape (figure 1(f)). The latter exhibits
tank-treading (TT) and oscillatory contractions similar
to the slippers seen by Fedosov et al. [4] (see the SI for
a movie and the insets in figure 5). Tank-treading refers
5
Figure 3. Experimental results: (a) Fraction of observed cell shapes as a function of the applied pressure drop (top axis) and
mean cell velocity (bottom axis). The horizontal error bars depict the standard deviation of the measured cell velocities for each
applied pressured drop. The shaded background is a guide to the eye. Furthermore, the insets show examples of experimental
images (see the SI for a collection of all photographs). (b) Estimated probability density function of the RBCs' center-of-mass
y-position within the channel for various pressure drops (indicated as numbers on the right in millibar) for all shapes combined.
We show the separated contributions of each shape to the distribution in the supplementary information. The area under the
curves is normalized to one. The dashed lines illustrate the wall positions. Both figures are for the position 10 mm downstream
from the channel entrance.
to the motion of the membrane around a (more or less)
static shape. Note that perfectly axisymmetric parachutes
are suppressed by the rectangular channel flow, contrary
to the situation for cylindrical tubes [4] or unbounded
Poiseuille flows [16].
To start the systematic study, we take a red blood cell
that is initially in the typical discocyte shape with its
rotation axis aligned along the tube's axis (cf. fig. 2). We
then vary the radial offset rinit from the center line as de-
scribed in section II B and record the final radial position
as well as the shape. The mean of the radial position is
extracted by a temporal average once the cell is in the
steady state (see the supplementary information for more
details). Figure 5 shows the result for a cell velocity of
≈ 6.5 mm/s. A single sharp transition at rinit ≈ 0.7 µm
from centered croissants to off-centered slippers is ob-
served. The final position of the slippers is mostly offset
only along the wider width of the channel (y-direction),
but not along the smaller height (z-direction). Hence we
find pronounced bistability: The result is significantly de-
termined by the initial condition and two different shapes
coexist. This is consistent with the 2D simulations by
Secomb et al. [19] and Tahiri et al. [6]. It also agrees
qualitatively with observations by Farutin and Misbah
for 3D simulations of vesicles in unbounded Poiseuille
flow [16].
To study the bistability in more detail, we vary the
imposed flow velocity as well as the initial offset rinit
Figure 4. Experimental results: Estimated probability density
function of the cells' center-of-mass y-position at the channel
entrance (position x = 0 mm). The pressure drops increase
from the bottom (20 mbar) to the top (1000 mbar) with the
numbers on the left side indicating the corresponding value
in millibar. The area under the curves is normalized to one.
The curves are offset in the vertical direction for illustration
purpose.
20501002003004005006007008009001000Appliedpressuredrop[mbar]01234567891000.20.40.60.810.920.830.190.430.530.470.380.250.150.160.140.130.250.360.540.70.850.850.850.870.080.170.80.520.210.170.08Measuredcellvelocity[mm/s]FractionofcellshapesCroissantsSlippersOthers(a)-6-4-2 0 2 4 620501002003004005006007008009001000(b)Probability density [a.u.]y-offset from center [µm]-6-4-2 0 2 4 620501002003004005006007008009001000Probability density [a.u.]y-offset from center [µm]6
an intermediate slipper state that can last several sec-
onds (see figure S4 and the movie in the supplementary
information). Hence, shapes observed after less than one
second often turn out to be transient, contrary to the in-
terpretation of Ye et al. [54] but in agreement with Prado
et al. [24].
Considering our results in figure 6(a) in more detail,
we find that two different types of croissants and slippers
are possible. On the one hand, at very low velocities
((cid:46) 0.7 mm/s) the slippers no longer exhibit tank-treading
motion of the membrane and instead show tumbling be-
havior: The cell rotates around the z-axis while approxi-
mately preserving its shape (similar to a rigid-body, see
the SI for a movie). The difference compared to the tum-
bling motion observed by Fedosov et al. [4] is that the cell
still exhibits a clear slipper-like instead of a proper disco-
cyte shape. Hence, we classify this mode still as slipper.
On the other hand, at very high velocities ((cid:38) 7 mm/s)
slightly asymmetric shapes strongly reminiscent of crois-
sants with a distinct tank-treading motion can sometimes
be observed (see the inset in figure 6(c) for an example).
As the shape itself is very close to a croissant, we will
nevertheless consider it to be a croissant below.
A natural question that occurs in light of the profound
bistability is the influence of other initial shapes on the
result. To this end, we consider a typical croissant as
well as a typical slipper as the starting shape. Both were
obtained from previous simulations that started with the
discocyte form and are depicted in the supplementary
information. We once again construct the shape phase
diagram as before and display the results in figures 6(b)
and (c). Note that the different starting shapes admit
a larger initial radial position rinit of the centroid. In
short, starting with a croissant favors croissants in the
steady state (the reddish area is larger than in figure 6(a)).
For slippers it is the other way around: Starting with
a slipper tends to produce more slippers (reddish area
smaller than in figure 6(a)). Despite this, the croissant-
only region from around 2 to 3 mm/s still exists unscathed.
Overall, only two qualitative differences occur between
the phase diagrams of different initial shapes, both at
lower velocity when starting with the croissant shape
(figure 6(b)): First, stable croissants emerge at very low
velocities ((cid:46) 0.7 mm/s) and second, the croissant-only
peak exhibits a "protrusion" into the slipper space. This
observation suggests that slippers and croissants can be
stable below 2 mm/s for most rinit values, although in
some cases a very precise croissant configuration is re-
quired in order to actually get a croissant in the steady
state.
Another interesting aspect concerns the radial posi-
tions of the centroids in the final steady states. The
average values are obtained by computing the tempo-
ral average in the steady state first for each simulation,
and then combining the results for identical shapes via a
weighted arithmetic mean. We use the observation time
in the steady state as the weight. This procedure leads
to figure 7(a). Obviously, the final radial positions are
Figure 5. Simulation results: Averaged radial position in the
steady state as a function of the initial radial offset for a
cell velocity of ≈ 6.5 mm/s. The RBC starts in the typical
discocyte shape with its rotation axis aligned with the tube's
axis (figure 2). The dotted line is a guide to the eye. Half of
the channel's extent along the y-direction (width) is shown
as a dashed line at the top. The extent in the z-direction
(height) is of less significance here since the steady states are
always almost centered in the z-direction. Furthermore, the
radial position for the converged slippers oscillates around a
mean value and their shapes show periodic "contractions" as
indicated by the vertical error bars and the right two insets,
respectively.
and characterize the behavior in the steady state. This
yields the shape phase diagram depicted in figure 6(a).
The cell velocity is extracted in the steady state via a
temporal average. For slippers the velocity varies peri-
odically (similar to the radial position): the minimum
and maximum in one period is indicated by the hor-
izontal error bars. Overall, the mean cell velocity u
ranges from 0.132 mm/s to 10.4 mm/s, matching with
the experimentally covered range. The corresponding
shear capillary number CaS := µu/κS varies therefore
in the interval CaS ∈ [0.0317, 2.50], while the bending
capillary number CaB := µuR2
RBC/κB lies in the range
CaB ∈ [8.45, 666]. The reddish area illustrates the approx-
imated region where croissants exist. Furthermore, there
is a maximal initial offset rinit above which overlapping
with the vessel wall would occur.
The shape phase diagram in figure 6(a) (together
with (b) and (c) explained below) constitutes our main
result from the simulations. Starting near the channel cen-
ter (in the reddish region) results in croissants, whereas
higher initial offsets lead to slippers. The transition is
found to be sharp, and depends significantly on the ve-
locity. Croissants are the only stable steady state in a
small region ranging from around 2 to 3 mm/s, indepen-
dently of the initial radial position. Smaller and larger
velocities tend to favor slippers. Stable croissants do not
appear below 0.25 mm/s. In the case of the slippers, the
final periodic state is usually reached after roughly 2 to
10 s. In contrast, the final croissant state is sometimes
achieved only after more than 10 to 30 s, possibly after
00.20.40.60.811.21.41.60246Initialradialpositionrinit[m]Steadystateradialposition[m]Radialpositiony-positionofchannelborder7
Figure 6. Simulation results: Shapes obtained when varying the initial offset rinit and the velocity. Each symbol corresponds to
one simulation. The horizontal axis shows the average cell velocity in the steady state, while horizontal error bars depict the
minimal and maximal velocities in one period (variations for croissants nearly zero and thus not visible). The upper dashed line
represents the maximal initial offset: Above this offset, the cell would overlap with the wall. The other lines and the colored
areas are guides to the eye and illustrate the different regions in the phase diagram. Each figure corresponds to a different initial
shape, namely (a) to the typical discocyte shape, (b) to a croissant and (c) to a slipper. These shapes are shown in figure 1(a)
and in the SI. The inset in the last figure depicts an example of a tank-treading croissant. Figure 5 corresponds to the vertical
column at ≈ 6.5 mm/s in sub-figure (a).
independent of the initial starting shape, i.e. a particular
steady state shape at a certain velocity is always located
at the same position. Furthermore, non-tank-treading
croissants are always almost centered, with only minor
deviations away from zero. These slight deviations in the
range from 2 to 4 mm/s are mainly due to some croissants
exhibiting minuscule periodic shape deformations. More-
over, the centroids of tank-treading croissants occurring
at velocities (cid:38) 8 mm/s are located near but not directly
in the center. Their slight off-centered position is a result
of their asymmetry.
In contrast to croissants, slippers are located 0.8 to
1.5 µm away from the channel's axis. The minimum po-
sition is attained for velocities near the border of the
croissant-only region in the phase diagram (at around 2
and 3 mm/s, compare figure 6). Above, the off-center po-
sition increases and seems to converge to a value of around
1.5 µm. The reason for this increase is that slippers be-
come more elongated and thinner at higher velocities (up
to a certain degree), as shown in figure 7(b) and also
observed in previous experiments [22]. Thus, they effec-
tively become smaller in the radial direction and their
centroids can move closer to the wall. We note that the
distance between the wall and the upper side of the slipper
approximately remains the same for all velocities. This
also hints at that the "optimal" off-center position for the
slippers is more than 1.5 µm away from the center, and
that this particular value is due to the smallness of the
channel.
V. COMPARISON BETWEEN EXPERIMENTS
AND SIMULATIONS
A. Comparison of shapes
Considering figure 1, the croissants obtained from sim-
ulations and experiments look very similar, although the
experimental shapes appear to be somewhat larger. The
reason is diffraction: The "true" cell border lies in the
bright and not within the dark rim. However, the slip-
pers appear to look qualitatively different. This is due
to the high magnification and numerical aperture of the
objective which results in a small depth of field of around
1 µm. Cell borders above and below the middle plane
are therefore blurred out and become invisible while the
mid-plane cut becomes dominant. Thus, for comparison
we should use the middle cross-section of the numerically
obtained shapes. Here we find good agreement (compare
figure 1(g) with (e)).
B. Comparison of the phase diagrams
A qualitative comparison between the phase diagrams
of steady states from the experiments (fig. 3(a)) and the
simulations (fig. 6) shows a striking resemblance: Both
exhibit a distinct peak in the number of croissants at
lower velocities (1 to 3 mm/s) at the expense of the num-
ber of slippers. The latter dominate the picture at high
velocities (> 7 mm/s). At intermediate velocities both
shapes coexist and can therefore be observed simulta-
neously in measurements. Moreover, the simulations at
very low velocities showed croissants only if the initial
(a)024681001234Cellvelocity[mm/s]Initialradialpositionrinit[m]Initializedasdiscocytenon-TTCroissantTTCroissantnon-TTSlipperTTSlipper(b)024681001234Cellvelocity[mm/s]Initializedascroissant(c)024681001234Cellvelocity[mm/s]Initializedasslipper8
Figure 7. Simulation results: (a) Average radial positions of the steady states from figure 6 as a function of cell velocity for the
three different starting shapes. The lower curves are for non-TT croissants and TT croissants, the upper curves are for (TT
and non-TT) slippers. We show on the vertical axis the weighted temporal mean of the radial centroid position of RBCs that
assume the same shapes. The vertical error bars depict the total minimal and maximal position, while the horizontal error bars
show the total minimal and maximal cell velocities (in each period of the steady states, respectively). (b) Extents of the slipper
shapes from figure (a) in the flow (x-)direction (length) and along the other two axes, as illustrated by the inset showing the
channel-aligned bounding box around a slipper. The vertical error bars depict the minimum and maximum extents during the
periodic contractions, while the horizontal error bars are the same as in (a).
RBC was already prepared in that state, meaning that in
the experiments this shape is highly unexpected. Indeed,
we were not able to clearly classify most of the observed
shapes in that regime as either croissants or slippers.
Obtaining a direct quantitative comparison requires a
translation of the numerical threshold in figure 6 (which
is in terms of the initial offset) into a prediction regarding
the fraction of shapes, because the experimental phase
diagram is in terms of the observed fraction of shapes.
This is done by counting the fraction of croissants entering
the channel with an offset below the numerical threshold.
This fraction corresponds directly to the predicted frac-
tion of croissant shapes. More precisely, we first define
rtrans as the initial radial offset which separates croissants
from slippers in the simulations by using the black line in
figure 6. An exception is the small croissant-only region
(i.e. the interval of the topmost horizontal line in figure 6)
where we take rtrans → ∞. This is consistent with our
interpretation that only croissants exist in this particular
interval. One rtrans is computed for each experimental
cell velocity from figure 3 (a). Second, each radial posi-
tion rtrans is projected onto the y-axis to give ytrans (see
sec. II B) because only the y-offset is known from experi-
ments. Third, from the experimental offset distribution
at the channel entrance (figure 4) we can then estimate
the fraction of cells φ that enter the channel with an
offset below ytrans. Accordingly, the simulations predict
a fraction φ of croissants in the steady state. The value
of φ can thus be directly compared with the experimental
phase diagram from figure 3 (a). This is done once for
every starting configuration employed in the simulations.
Figure 8 shows this key result of our contribution, i.e.
the predicted fraction of croissants φ as a function of
the cell velocity for each starting shape. The vertical
error bars depict the uncertainty in the prediction, whose
computation is explained in the supplementary informa-
tion. They are comparably large in the croissant-only
region because the experimental velocities lie very near
its sharp boundary. The horizontal error bars illustrate
the standard deviation σu of the experimentally mea-
sured cell velocities. Clearly, we find very good agreement
between the prediction from the simulation and the exper-
imental observation when considering the slipper starting
shape (figure 8(c)). Starting with a discocyte or croissant
leads to slightly more pronounced deviations (figures 8(a)
and (b)), but still a satisfactory semi-quantitative agree-
ment is maintained. This suggests the intuitive conclusion
that the starting shapes in the experiment are closer to
the rather asymmetric slippers than to the highly symmet-
ric discocytes or croissants. Indeed, as explicitly shown
in the SI, we only observe non-classifiable and rather
asymmetric "other" shapes at the channel entrance.
As mentioned in the introduction, experimental investi-
gations with more detailed shape studies are rather scarce.
A comparison of the phase diagrams with the experimental
literature is therefore limited to rough qualitative state-
ments. Tomaiuolo et al. [22] found croissants and "others"
for a cell velocity of 1.1 mm/s using λ ≈ 5 in a cylin-
drical tube with diameter 10 µm. This is in agreement
with our results. At 36 mm/s, slippers but also croissants
have been observed. Since we cannot reach velocities that
high, we can neither confirm nor refute the occurrence of
the latter. Extrapolation of figure 8 is dangerous since
the Reynolds number at 36 mm/s is around Re ≈ 0.24
024681000.511.52SlippersSlippersTTCroiss.non-TTCroissants(a)Cellvelocity[mm/s]Averagesteadystateradialposition[m]Init.asdiscocyteInit.ascroissantInit.asslippery-extentz-extentNoslippers(b)0246810246810Cellvelocity[mm/s]RBCy-andz-extents[m]Lengthz-extenty-extentLength46810RBClength[m]9
Figure 8. Fraction of croissants φ predicted by the simulations, once for each starting configuration employed in the simulations:
(a) Simulations started with the typical discocyte, (b) with the croissant and (c) with the slipper shape. To allow for a direct
comparison, we included the experimental results from figure 3(a) in each diagram (black dashed line). The horizontal error
bars depict the standard deviation σu of the measured cell velocities (as in figure 3(a)), while the vertical error bars show the
uncertainty in the prediction as explained in the supplementary information. The lines and shaded areas serve as guides to the
eye. See the main text for further details.
and thus inertia effects might have noticeable contribu-
tions [89, 90]. Continuing, Cluitmans et al. [25] found
croissants and tumbling "others" at 1.1 mm/s and slippers
at 13.6 mm/s in rectangular channels of 10 µm and 7 µm
widths and a height of 10 µm, which is consistent with our
results. The experimental phase diagram presented in ref-
erences 21 and 20 also agrees with our results insofar that
slippers occur at higher and croissants at lower velocities.
Yet, the considered velocities were higher than 10 mm/s
and the viscosity ratio was λ (cid:46) 0.27, i.e. much lower.
Furthermore, figure 3 in reference 18 (cylindrical tube,
λ ≈ 4) also showed coexistence of croissants and slippers
for velocities (cid:46) 1 mm/s and only croissants roughly in
the range 1 – 2 mm/s, matching approximately with our
results.
Regarding previous numerical studies, Fedosov et al. [4]
performed detailed 3D numerical simulations in cylindri-
cal channels for λ = 1. Taking a diameter of 10 µm
(translating into a confinement value of χ = 0.65 in
their work), they varied the average velocity from around
0.05 mm/s to 0.7 mm/s. They observed a transition from
snaking, to tumbling, to tank-treading slippers and finally
to parachutes (which are very similar to croissants). In
our simulations we found tumbling and tank-treading
slippers at velocities of the order of 0.1 mm/s, and an
increasing frequency of croissants above. This matches at
least qualitatively with Fedosov et al.'s results. However,
they did not vary the initial condition.
C. Comparison of cell positions
Next, we compare the preferred position of the cells
in the steady state. The simulations predict a centered
positioning of croissants (figure 7(a)), i.e. both the y- and
the z-offsets are nearly zero. This matches with figure 3(b)
Figure 9. Comparison between the centroid positions from the
simulations (absolute values of the y- and z-coordinates) and
experiments (absolute value of the y-coordinate) for cells that
have a TT-slipper shape in the steady state. Error bars for
the simulations as in figure 7(a). The horizontal error bars
for the experimental data depict the standard deviation σu of
the cell velocities, while the vertical error bars represent the
estimated uncertainty in the position determination.
where a very sharp peak at the channel center is found
for the pressure drops within the croissant-peak region.
For slippers, the simulations showed an increase of the
radial position of up to around 1.5 µm (figure 7(a)). Con-
sidering the y- and z-coordinates separately in figure 9,
we see that z ≈ 0 and the major offset happens in the
y-direction. This is rather fortunate as the y-offset is
also easily accessible in the experiments, contrary to the
z-offset. As can be seen in the measured y-distribution
(figure 3(b)), we have two off-centered peaks for slippers.
Taking the distribution function for only the slippers, we
extract the positions yl and yr of the two peaks. Ex-
ploiting the ±y-symmetry of the channel, the off-centered
position is then computed as (yr − yl)/2, i.e. in essence as
024681000.20.40.60.81(a)Cellvelocity[mm/s]FractionofcroissantsφInitializedasdiscocyteInit.asdiscocyteExperiment024681000.20.40.60.81(b)Cellvelocity[mm/s]InitializedascroissantInit.ascroissantExperiment024681000.20.40.60.81(c)Cellvelocity[mm/s]InitializedasslipperInit.asslipperExperiment24681000.511.52Cellvelocity[mm/s]Position[m]Simulation:ySimulation:zExperiment:ythe average of the two peak distances to the central mini-
mum. Figure 9 compares these values with the numerical
results: The behavior is the same (an increase with veloc-
ity) and the predicted values show only a small systematic
deviation of around ≈ 0.3 µm, i.e. of less than 4 % of the
RBC diameter 2RRBC. A possible reason is that the op-
tically recorded boundaries of the RBC and the channel
walls are somewhat blurry (compare the experimental
images in figure 1).
D.
Implications of the comparison
There has been quite some debate in the literature if
the croissant (or parachute) shapes observed via light
microscopy are indeed what they appear to be. Gaeht-
gens et al. [17] (fig. 4 therein), for example, solidified the
flowing RBCs with glutaraldehyde and found that the
croissant-like shapes were actually slipper-like. Skalak
and Branemark [37] pointed out that such shapes can also
be "edge-on" discocytes with a flattened back. Ultimately,
to uniquely identify the forms one needs some method
to record the full 3D geometry of the flowing cells (e.g.
as in references 15, 26, 32, 91–94). This is unfortunately
very hard to implement in the present experimental setup.
However, this missing information is complemented here
by the numerical simulations which are in good agreement
with the experiments and thus our interpretation of the
shapes as croissants should be correct.
The good agreement furthermore implies that our red
blood cell model and simulation method is fully appropri-
ate for describing the flow of RBCs in a straight microchan-
nel. More sophisticated methods including e.g. thermal
fluctuations or surface viscosity [4, 24, 52, 53, 95, 96] are,
at least for the present geometry, not required. For crois-
sants this is intuitive since membrane movement such as
tank-treading is absent, for the tank-treading slippers it
is somewhat less obvious.
VI. SUMMARY & CONCLUSION
To summarize, we have performed in-vitro experiments
and 3D simulations of healthy red blood cells flowing in
a microchannel. The viscosity ratio was approximately
5 and the flow velocities ranged from around 0.1 mm/s
10
to 10 mm/s in both methodologies, corresponding to the
typical conditions prevailing in the microvascular system.
We found that both the flow velocity as well as the initial
starting configuration (offset from channel center, shape)
have a major impact on the final steady state of the cells.
Using three different starting shapes (discocyte, croissant,
slipper), we constructed the corresponding phase diagrams
via simulations. In most cases the cells assumed one out
of two different forms: either a centered croissant or
an off-centered slipper. Interestingly, for most velocities
bistability, i.e. a dependence of the final shape on the
initial position, was observed. Only in a small range
of velocities (at around ≈ 1 mm/s) was the final shape
found to be always a croissant. The experimental diagram
showed very good agreement with the numerical result,
especially when considering the simulations that used the
rather asymmetric slipper as starting shape.
We thus conclude that the employed numerical RBC
model can sensibly describe the cell behavior in the pre-
sented setup. Moreover, since we used physiological viscos-
ity ratios and flow velocities, we speculate that croissants
and slippers can occur in the microvasculature at the
same set of system parameters not just as transients but
rather that both are states which are intrinsically assumed
by the cells. Our results are important for applications
where the cells should be in a specific state (e.g. in lab-on-
a-chip devices) and allow for a comprehensive validation
of numerical models.
CONFLICTS OF INTEREST
There are no conflicts to declare.
ACKNOWLEDGMENTS
A. Guckenberger and A. Kihm contributed equally to
this work. S. Gekle and C. Wagner contributed equally
to this work.
Funding from the Volkswagen Foundation and com-
puting time granted by the Leibniz-Rechenzentrum on
SuperMUC are gratefully acknowledged by A. Gucken-
berger and S. Gekle. A. Kihm, T. John and C. Wag-
ner kindly acknowledge the support and funding of the
"Deutsch-Französische-Hochschule" (DFH) DFDK "Living
Fluids".
[1] J. B. Freund, "The flow of red blood cells through a narrow
spleen-like slit," Phys. Fluids 25, 110807 (2013).
[2] J. Picot, P. A. Ndour, S. D. Lefevre, W. El Nemer,
H. Tawfik, J. Galimand, L. Da Costa, J.-A. Ribeil, M. de
Montalembert, V. Brousse, B. Le Pioufle, P. Buffet,
C. Le Van Kim, and O. Français, "A biomimetic mi-
crofluidic chip to study the circulation and mechanical
retention of red blood cells in the spleen," Am. J. Hematol.
90, 339 (2015).
[3] S. Salehyar and Q. Zhu, "Deformation and internal stress
in a red blood cell as it is driven through a slit by an
incoming flow," Soft Matter 12, 3156 (2016).
[4] D. A. Fedosov, M. Peltomäki, and G. Gompper, "Defor-
mation and dynamics of red blood cells in flow through
cylindrical microchannels," Soft Matter 10, 4258 (2014).
[5] O. Aouane, M. Thiébaud, A. Benyoussef, C. Wagner,
and C. Misbah, "Vesicle dynamics in a confined Poiseuille
flow: From steady state to chaos," Phys. Rev. E 90, 033011
(2014).
[6] N. Tahiri, T. Biben, H. Ez-Zahraouy, A. Benyoussef, and
C. Misbah, "On the problem of slipper shapes of red blood
cells in the microvasculature," Microvasc. Res. 85, 40
(2013).
[7] V. Vitkova, M.-A. Mader, B. Polack, C. Misbah, and
T. Podgorski, "Micro-Macro Link in Rheology of Ery-
throcyte and Vesicle Suspensions," Biophys. J. 95, L33
(2008).
[8] D. A. Fedosov, W. Pan, B. Caswell, G. Gompper, and
G. E. Karniadakis, "Predicting human blood viscosity in
silico," Proc. Natl. Acad. Sci. 108, 11772 (2011).
[9] T. Krüger, M. Gross, D. Raabe,
and F. Varnik,
"Crossover from tumbling to tank-treading-like motion
in dense simulated suspensions of red blood cells," Soft
Matter 9, 9008 (2013).
[10] M. Thiébaud, Z. Shen, J. Harting, and C. Misbah, "Pre-
diction of Anomalous Blood Viscosity in Confined Shear
Flow," Phys. Rev. Lett. 112, 238304 (2014).
[11] D. Katanov, G. Gompper, and D. A. Fedosov, "Microvas-
cular blood flow resistance: Role of red blood cell migration
and dispersion," Microvasc. Res. 99, 57 (2015).
[12] L. Lanotte, J. Mauer, S. Mendez, D. A. Fedosov, J.-M.
Fromental, V. Claveria, F. Nicoud, G. Gompper, and
M. Abkarian, "Red cells' dynamic morphologies govern
blood shear thinning under microcirculatory flow condi-
tions," Proc. Natl. Acad. Sci. 113, 13289 (2016).
[13] E. Henry, S. H. Holm, Z. Zhang, J. P. Beech, J. O. Tegen-
feldt, D. A. Fedosov, and G. Gompper, "Sorting cells by
their dynamical properties," Sci. Rep. 6, 34375 (2016).
[14] O. Otto, P. Rosendahl, A. Mietke, S. Golfier, C. Herold,
D. Klaue, S. Girardo, S. Pagliara, A. Ekpenyong, A. Ja-
cobi, M. Wobus, N. Töpfner, U. F. Keyser, J. Mansfeld,
E. Fischer-Friedrich, and J. Guck, "Real-time deforma-
bility cytometry: On-the-fly cell mechanical phenotyping,"
Nat. Methods 12, 199 (2015).
[15] F. Merola, P. Memmolo, L. Miccio, R. Savoia, M. Mug-
nano, A. Fontana, G. D'Ippolito, A. Sardo, A. Iolascon,
A. Gambale, and P. Ferraro, "Tomographic Flow Cytom-
etry by Digital Holography," Light Sci. Appl. 6, e16241
(2017).
[16] A. Farutin and C. Misbah, "Symmetry breaking and cross-
streamline migration of three-dimensional vesicles in an
axial Poiseuille flow," Phys. Rev. E 89, 042709 (2014).
[17] P. Gaehtgens, C. Dührssen, and K. H. Albrecht, "Motion,
deformation, and interaction of blood cells and plasma
during flow through narrow capillary tubes," Blood Cells
6, 799 (1980).
[18] Y. Suzuki, N. Tateishi, M. Soutani, and N. Maeda, "De-
formation of Erythrocytes in Microvessels and Glass Cap-
illaries: Effects of Erythrocyte Deformability," Microcir-
culation 3, 49 (1996).
[19] T. W. Secomb, B. Styp-Rekowska, and A. R. Pries, "Two-
Dimensional Simulation of Red Blood Cell Deformation
and Lateral Migration in Microvessels," Ann. Biomed.
Eng. 35, 755 (2007).
[20] M. Faivre, Drops, Vesicles and Red Blood Cells: Deforma-
bility and Behavior under Flow, Ph.D. thesis, Université
Joseph-Fourier - Grenoble I, Grenoble (2006).
[21] M. Abkarian, M. Faivre, R. Horton, K. Smistrup, C. A.
Best-Popescu, and H. A. Stone, "Cellular-scale hydrody-
11
namics," Biomed. Mater. 3, 034011 (2008).
[22] G. Tomaiuolo, M. Simeone, V. Martinelli, B. Rotoli, and
S. Guido, "Red blood cell deformation in microconfined
flow," Soft Matter 5, 3736 (2009).
[23] G. Tomaiuolo and S. Guido, "Start-up shape dynamics of
red blood cells in microcapillary flow," Microvasc. Res. 82,
35 (2011).
[24] G. Prado, A. Farutin, C. Misbah, and L. Bureau, "Vis-
coelastic Transient of Confined Red Blood Cells," Biophys.
J. 108, 2126 (2015).
[25] J. C. A. Cluitmans, V. Chokkalingam, A. M. Janssen,
R. Brock, W. T. S. Huck, and G. J. C. G. M. Bosman,
"Alterations in Red Blood Cell Deformability during Stor-
age: A Microfluidic Approach," BioMed Res. Int. 2014,
e764268 (2014).
[26] S. Quint, A. F. Christ, A. Guckenberger, S. Himbert,
L. Kaestner, S. Gekle, and C. Wagner, "3D tomography
of cells in micro-channels," Appl. Phys. Lett. 111, 103701
(2017).
[27] R. M. Hochmuth, R. N. Marple, and S. P. Sutera, "Cap-
illary blood flow: I. Erythrocyte deformation in glass cap-
illaries," Microvascular Research 2, 409 (1970).
[28] V. Seshadri, R. M. Hochmuth, P. A. Croce, and S. P.
Sutera, "Capillary blood flow: III. Deformable model cells
compared to erythrocytes in vitro," Microvascular Research
2, 434 (1970).
[29] V. P. Zharov, E. I. Galanzha, Y. Menyaev, and V. V.
Tuchin, "In vivo high-speed imaging of individual cells in
fast blood flow," J. Biomed. Opt 11, 054034 (2006).
[30] G. Tomaiuolo, V. Preziosi, M. Simeone, S. Guido, R. Cian-
cia, V. Martinelli, C. Rinaldi, and B. Rotoli, "A method-
ology to study the deformability of red blood cells flowing
in microcapillaries in vitro," Ann. Ist. Super. Sanità 43,
186 (2007).
[31] S. Guido and G. Tomaiuolo, "Microconfined flow behavior
of red blood cells in vitro," Comptes Rendus Phys. 10, 751
(2009).
[32] S. S. Gorthi and E. Schonbrun, "Phase imaging flow
cytometry using a focus-stack collecting microscope," Opt.
Lett. 37, 707 (2012).
[33] L. Lanotte, G. Tomaiuolo, C. Misbah, L. Bureau, and
S. Guido, "Red blood cell dynamics in polymer brush-
coated microcapillaries: A model of endothelial glycocalyx
in vitro," Biomicrofluidics 8, 014104 (2014).
[34] G. Tomaiuolo, L. Lanotte, R. D'Apolito, A. Cassinese,
and S. Guido, "Microconfined flow behavior of red blood
cells," Med. Eng. Phys. 38, 11 (2016).
[35] V. Clavería, O. Aouane, M. Thiébaud, M. Abkarian,
G. Coupier, C. Misbah, T. John, and C. Wagner, "Clus-
ters of red blood cells in microcapillary flow: Hydrodynamic
versus macromolecule induced interaction," Soft Matter
12, 8235 (2016).
[36] M. M. Guest, T. P. Bond, R. G. Cooper, and J. R. Derrick,
"Red Blood Cells: Change in Shape in Capillaries," Science
142, 1319 (1963).
[37] R. Skalak and P. I. Branemark, "Deformation of Red
Blood Cells in Capillaries," Science 164, 717 (1969).
[38] K. Kubota, J. Tamura, T. Shirakura, M. K Imura,
K. Y Amanaka, T. Isozaki, and I. Nishio, "The behaviour
of red cells in narrow tubes in vitro as a model of the
microcirculation," Br. J. Haematol. 94, 266 (1996).
[39] G. Tomaiuolo, L. Lanotte, G. Ghigliotti, C. Misbah, and
S. Guido, "Red blood cell clustering in Poiseuille micro-
capillary flow," Phys. Fluids 24, 051903 (2012).
[40] C. Wagner, P. Steffen, and S. Svetina, "Aggregation of
red blood cells: From rouleaux to clot formation," Comptes
Rendus Physique Living fluids / Fluides vivants, 14, 459
(2013).
[41] M. Brust, O. Aouane, M. Thiébaud, D. Flormann,
C. Verdier, L. Kaestner, M. W. Laschke, H. Selmi,
A. Benyoussef, T. Podgorski, G. Coupier, C. Misbah,
and C. Wagner, "The plasma protein fibrinogen stabilizes
clusters of red blood cells in microcapillary flows," Sci.
Rep. 4, 4348 (2014).
[42] H. L. Goldsmith and J. Marlow, "Flow Behaviour of Ery-
throcytes. I. Rotation and Deformation in Dilute Suspen-
sions," Proc. R. Soc. Lond. B Biol. Sci. 182, 351 (1972).
[43] T. W. Secomb, R. Skalak, N. Özkaya, and J. F. Gross,
"Flow of axisymmetric red blood cells in narrow capillar-
ies," J. Fluid Mech. 163, 405 (1986).
[44] T. W. Secomb, "Flow-dependent rheological properties
of blood in capillaries," Microvascular Research 34, 46
(1987).
[45] T. W. Secomb, R. Hsu, and A. R. Pries, "Motion of
red blood cells in a capillary with an endothelial surface
layer: Effect of flow velocity," Am. J. Physiol. - Heart
Circ. Physiol. 281, H629 (2001).
[46] T. W. Secomb and R. Skalak, "A two-dimensional model
for capillary flow of an asymmetric cell," Microvascular
Research 24, 194 (1982).
[47] B. Kaoui, G. Biros, and C. Misbah, "Why Do Red Blood
Cells Have Asymmetric Shapes Even in a Symmetric
Flow?" Phys. Rev. Lett. 103, 188101 (2009).
[48] B. Kaoui, N. Tahiri, T. Biben, H. Ez-Zahraouy, A. Beny-
oussef, G. Biros, and C. Misbah, "Complexity of vesicle
microcirculation," Phys. Rev. E 84, 041906 (2011).
[49] B. Kaoui, T. Krüger, and J. Harting, "How does confine-
ment affect the dynamics of viscous vesicles and red blood
cells?" Soft Matter 8, 9246 (2012).
[50] L. Shi, T.-W. Pan, and R. Glowinski, "Deformation of
a single red blood cell in bounded Poiseuille flows," Phys.
Rev. E 85, 016307 (2012), 10.1103/PhysRevE.85.016307.
[51] G. R. Lázaro, A. Hernández-Machado, and I. Pagonabar-
raga, "Rheology of red blood cells under flow in highly
confined microchannels: I. effect of elasticity," Soft Mat-
ter 10, 7195 (2014).
[52] H. Noguchi and G. Gompper, "Shape transitions of fluid
vesicles and red blood cells in capillary flows," Proc. Natl.
Acad. Sci. 102, 14159 (2005).
[53] J. L. McWhirter, H. Noguchi, and G. Gompper, "Defor-
mation and clustering of red blood cells in microcapillary
flows," Soft Matter 7, 10967 (2011).
[54] T. Ye, H. Shi, L. Peng, and Y. Li, "Numerical studies
of a red blood cell in rectangular microchannels," J. Appl.
Phys. 122, 084701 (2017).
[55] B. Kaoui, G. Coupier, C. Misbah, and T. Podgorski,
"Lateral migration of vesicles in microchannels: Effects of
walls and shear gradient," Houille Blanche , 112 (2009).
[56] D. Cordasco, A. Yazdani, and P. Bagchi, "Comparison of
erythrocyte dynamics in shear flow under different stress-
free configurations," Phys. Fluids 26, 041902 (2014).
[57] Z. Peng, A. Mashayekh, and Q. Zhu, "Erythrocyte re-
sponses in low-shear-rate flows: Effects of non-biconcave
stress-free state in the cytoskeleton," J. Fluid Mech. 742,
96 (2014).
[58] K. Sinha and M. D. Graham, "Dynamics of a single red
blood cell in simple shear flow," Phys. Rev. E 92, 042710
(2015).
12
[59] G. R. Cokelet and H. J. Meiselman, "Rheological Com-
parison of Hemoglobin Solutions and Erythrocyte Suspen-
sions," Science 162, 275 (1968).
[60] A. S. Popel and P. C. Johnson, "Microcirculation and
Hemorheology," Annu. Rev. Fluid Mech. 37, 43 (2005).
[61] B. Cui, H. Diamant, and B. Lin, "Screened Hydrodynamic
Interaction in a Narrow Channel," Phys. Rev. Lett. 89,
188302 (2002).
[62] H. Diamant, "Hydrodynamic Interaction in Confined Ge-
ometries," J. Phys. Soc. Jpn. 78, 041002 (2009).
[63] A. Koller, B. Dawant, A. Liu, A. S. Popel, and P. C.
Johnson, "Quantitative analysis of arteriolar network ar-
chitecture in cat sartorius muscle," Am. J. Physiol. - Heart
Circ. Physiol. 253, H154 (1987).
[64] E. Evans and Y.-C. Fung, "Improved measurements of the
erythrocyte geometry," Microvasc. Res. 4, 335 (1972).
[65] D.-V. Le, "Subdivision elements for large deformation of
liquid capsules enclosed by thin shells," Comput. Methods
Appl. Mech. Eng. 199, 2622 (2010).
[66] S. Chien, S. Usami, H. M. Taylor, J. L. Lundberg, and
M. I. Gregersen, "Effects of hematocrit and plasma pro-
teins on human blood rheology at low shear rates." J. Appl.
Physiol. 21, 81 (1966).
[67] R. Skalak, N. Ozkaya, and T. C. Skalak, "Biofluid Me-
chanics," Annu. Rev. Fluid Mech. 21, 167 (1989).
[68] T. W. Secomb, "Blood Flow in the Microcirculation,"
Annu. Rev. Fluid Mech. 49, 443 (2017).
[69] Y. Kim, H. Shim, K. Kim, H. Park, S. Jang, and Y. Park,
"Profiling individual human red blood cells using common-
path diffraction optical tomography," Sci. Rep. 4, 6659
(2014).
[70] R. Skalak, A. Tozeren, R. P. Zarda, and S. Chien, "Strain
Energy Function of Red Blood Cell Membranes," Biophys.
J. 13, 245 (1973).
[71] T. Krüger, F. Varnik, and D. Raabe, "Efficient and
accurate simulations of deformable particles immersed
in a fluid using a combined immersed boundary lattice
Boltzmann finite element method," Comput. Math. Appl.
61, 3485 (2011).
[72] Y.-Z. Yoon, J. Kotar, G. Yoon, and P. Cicuta, "The
nonlinear mechanical response of the red blood cell," Phys.
Biol. 5, 036007 (2008).
[73] J. B. Freund, "Numerical Simulation of Flowing Blood
Cells," Annu. Rev. Fluid Mech. 46, 67 (2014).
[74] P. B. Canham, "The minimum energy of bending as a
possible explanation of the biconcave shape of the human
red blood cell," J. Theor. Biol. 26, 61 (1970).
[75] W. Helfrich, "Elastic Properties of Lipid Bilayers: Theory
and Possible Experiments," Z. Naturforsch. C 28, 693
(1973).
[76] A. Guckenberger and S. Gekle, "Theory and algorithms
to compute Helfrich bending forces: A review," J. Phys.
Condens. Matter 29, 203001 (2017).
[77] Y. Park, C. A. Best, K. Badizadegan, R. R. Dasari, M. S.
Feld, T. Kuriabova, M. L. Henle, A. J. Levine, and
G. Popescu, "Measurement of red blood cell mechanics
during morphological changes," Proc. Natl. Acad. Sci. 107,
6731 (2010).
[78] A. Guckenberger, M. P. Schraml, P. G. Chen, M. Leonetti,
and S. Gekle, "On the bending algorithms for soft objects
in flows," Comput. Phys. Comm. 207, 1 (2016).
[79] A. Farutin, T. Biben, and C. Misbah, "3D numerical
simulations of vesicle and inextensible capsule dynamics,"
J. Comput. Phys. 275, 539 (2014).
13
[80] A. Guckenberger and S. Gekle, "A boundary integral
method with volume-changing objects for ultrasound-
triggered margination of microbubbles," arXiv:1608.05196
[physics] (2017), arXiv:1608.05196 [physics].
[81] C. Pozrikidis, "Interfacial Dynamics for Stokes Flow," J.
Comput. Phys. 169, 250 (2001).
[82] H. Zhao, A. H. Isfahani, L. N. Olson, and J. B. Freund,
"A spectral boundary integral method for flowing blood
cells," J. Comput. Phys. 229, 3726 (2010).
[83] D. Saintillan, E. Darve, and E. S. G. Shaqfeh, "A smooth
particle-mesh Ewald algorithm for Stokes suspension sim-
ulations: The sedimentation of fibers," Phys. Fluids 17,
033301 (2005).
[84] Y. Saad and M. Schultz, "GMRES: A Generalized Mini-
mal Residual Algorithm for Solving Nonsymmetric Linear
Systems," SIAM J. Sci. Stat. Comput. 7, 856 (1986).
[85] P. Bogacki and L. F. Shampine, "A 3(2) pair of Runge -
Kutta formulas," Appl. Math. Lett. 2, 321 (1989).
[86] A. Daddi-Moussa-Ider, A. Guckenberger, and S. Gekle,
"Long-lived anomalous thermal diffusion induced by elastic
cell membranes on nearby particles," Phys. Rev. E 93,
012612 (2016).
[87] A. R. Pries, T. W. Secomb, and P. Gaehtgens, "Structure
and hemodynamics of microvascular networks: Hetero-
geneity and correlations," Am. J. Physiol. - Heart Circ.
Physiol. 269, H1713 (1995).
[88] O. Baskurt, B. Neu, and H. Meiselman, Red Blood Cell
Aggregation (CRC Press, 2011).
[89] B. Kaoui and J. Harting, "Two-dimensional lattice Boltz-
mann simulations of vesicles with viscosity contrast,"
Rheol Acta 55, 465 (2016).
[90] C. Schaaf and H. Stark, "Inertial migration and axial
control of deformable capsules," Soft Matter 13, 3544
(2017).
[91] N. C. Pégard and J. W. Fleischer, "Three-dimensional
deconvolution microfluidic microscopy using a tilted chan-
nel," J. Biomed. Opt. 18, 040503 (2013).
[92] N. C. Pégard, M. L. Toth, M. Driscoll, and J. W. Fleis-
cher, "Flow-scanning optical tomography," Lab Chip 14,
4447 (2014).
[93] V. K. Jagannadh, M. D. Mackenzie, P. Pal, A. K. Kar,
and S. S. Gorthi, "Slanted channel microfluidic chip for
3D fluorescence imaging of cells in flow," Opt. Express
24, 22144 (2016).
[94] K. Kim, K. Choe, I. Park, P. Kim, and Y. Park, "Holo-
graphic intravital microscopy for 2-D and 3-D imaging
intact circulating blood cells in microcapillaries of live
mice," Sci. Rep. 6, 33084 (2016).
[95] G. Tomaiuolo, M. Barra, V. Preziosi, A. Cassinese, B. Ro-
toli, and S. Guido, "Microfluidics analysis of red blood
cell membrane viscoelasticity," Lab Chip 11, 449 (2011).
[96] A. Yazdani and P. Bagchi, "Influence of membrane vis-
cosity on capsule dynamics in shear flow," J. Fluid Mech.
718, 569 (2013).
Supplementaryinformationfor"Numerical-experimentalobservationofshapebistabilityofredbloodcellsflowinginamicrochannel"AchimGuckenbergera,AlexanderKihmb,ThomasJohn,bChristianWagner,bcStephanGekleaDated:November19,2017ContentsS1Croissantandslipperinitialshapes1S2Timeevolutiondetails2S3Abouttheerrorbarsintheprediction3S4Additionalexperimentaldata5S5References5S6Rawexperimentalimagesatx=10mm7S7Rawexperimentalimagesatx=0mm12S1CroissantandslipperinitialshapesFigureS1showstheemployedredbloodcell(RBC)shapeswhentheinitialshapeistakentobeacroissantorslipper.Theseshapeswereobtainedfromaprevioussimulationwherewestartedwiththetypicaldiscocyteshape.FigureS2depictsthecorrespondingsimulationsetups,whichareidenticaltotheonefromthemaintextexceptforthedifferentRBCshape.Especiallynotethattheinitialradialoffsetrinitofthecentroidisalongthesameline.Wealwaysusethesamecroissantorslippershape,regardlessofthevalueofrinit.(a)(b)(c)FigureS1:Employedinitialshapesinthesimulationswhenstartingasa(a)croissantor(b)slipper.Figure(c)showsthecross-sectionoftheslipperfrom(b).Theblacklinesonthesurfacesrepresenttheusedmesh.aBiofluidSimulationandModeling,FachbereichPhysik,UniversitätBayreuth,BayreuthbExperimentalPhysics,SaarlandUniversity,66123,Saarbrücken,GermanycPhysicsandMaterialsScienceResearchUnit,UniversityofLuxembourg,Luxembourg,Luxembourg1FigureS2:Thesimulationsetupswhenstartingwithacroissant(left)oraslippershape(right),similartofigure2fromthemaintext.ThecellshapesaretheonesfromfigureS1.S2TimeevolutiondetailsAsnotedinthemaintextandshownintheTTSlippersupplementaryvideo,thetank-treading(TT)slipperexhibitsoscillatorycontractions.Theseresultinperiodicvariationsoftheradialpositionandthecellvelocity,asexemplifiedinfigureS3.Thisfigurealsoillustrateshowweextracttheaverage,minimalandmaximalvaluesafterreachingthesteadystate.Thesimulationresultsfromthemaintextdepicttheaveragevaluesasthemaindatapointsandtheminimalandmaximalvaluesviaerrorbars.Notethatwedothesamefortheothershapes,althoughtheresultingerrorbarsaretoosmalltobeseeninthefigures.01234560246(a)Time[s]Radialposition[m]Min./avg./max.radialpositionRadialpositiony-positionofchannelborder01234565.566.57(b)Time[s]Cellvelocity[mm/s]Min./avg./max.cellvelocityCellvelocityFigureS3:Timeevolutionof(a)theradialpositionand(b)thecellvelocityforaslippershape.Thedataisforthenumericalsimulationwithrinit=1.2µmfromfigure5inthemaintext.Theorangelinesshowfrombottomtotoptheminimum,averageandmaximumvaluesthatareextractedinthesteadystate,whichistakentobeginat2sinthisparticularexample.Furthermore,reachingthesteadystateoftentakesafewseconds.Convergenceintothecroissantshapeusuallytakeslongerthanreachingasteadyslipperstate.ThisisillustratedinfigureS4(a)whereweshowforeachsimulationfromfigure6inthemaintexttheapproximatedtimeuntilthesteadystateisreached.Thistimeisthedurationmeasuredfromthestartofthesimulationsuntiltheposition,shape,velocityanddeformationofthecellsnolongerchangeorbecomeperiodic.Thelongesttimesareobservedwhenthevelocityliesinthecroissant-onlyrange.AnexampleforsuchacaseisdisplayedinfigureS4(b):Foraroundeightseconds,thecellisinanalmostperiodicslipperstatebeforemovingtothecenterandbecomingacroissant.However,afteranotherfoursecondssomemembranerotationoccurs,i.e.theRBCdimples(whicharespecialpointsduetothediscocytereferencestate)movetoaslightlydifferentlocation.Thisresultsinashortlivedandslightlyoff-centeredposition.Afteratotaltimeofaround14sthecellisinthefinalcroissantstate,withnomovementoccurringanymore.SeethevideoLongCroissantforanillustration.202468100102030(a)Cellvelocity[mm/s]Approx.timeuntilsteadystate[s]non-TTC.TTC.non-TTS.TTS.Init.asdiscocyteInit.ascroissantInit.asslipper02468101214160246(b)IntermediateslipperstateMembranerotationIntermediatecroissantFinalcroissantTime[s]Radialposition[m]Radialpositiony-positionofchannelborderFigureS4:(a)Estimatesofthetimeittakesforthecellstoreachthesteadystate.Thefigureshowsthistimeforallsimulationsfromthethreediagramsfromfigure6inthemaintext,wherethecellwasinitializedasdiscocyte,croissantandslipper,respectively.Thesethreeinitialshapesareindicatedbythepurple,orangeandgreenbordersaroundthesymbols(seethethreerowsinthelegend).Thefinalsteadystateshapes(non-tank-treadingcroissant,tank-treadingcroissant,non-tank-treadingslipperandtank-treadingslipper)arerepresentedbythesamesymbolsasinthemaintext(comparethefourcolumnsinthelegend).(b)Timeevolutionoftheradialpositionofacellwhichisinitiallyinthediscocytestatewithrinit≈1.89µmandhasanaveragevelocityof≈2.79mm/sinthesteadystate(i.e.itliesinthecroissant-onlyregion).Thesteadystatebeginsataround14s.SeethemovieLongCroissantfora3Dvisualization.S3AbouttheerrorbarsinthepredictionThedeterminationoftheverticalerrorbarsinthecomparisonbetweenexperimentsandsimulations(figure8inthemaintext)consistsofseveralstepsthatwillbedescribedinthefollowing.Tothisend,considerfigureS5.Thisfigureshowsexemplarilythenumericalphasediagramwhenthestartingshapeisthediscocyte,i.e.thesymbolsthatindicatethesteadystatesareidenticaltofigure6(a)fromthemaintext.Themiddlegraylinerepresentsthepositionoftheapproximatedtransitionthresholdrtransbetweencroissantsandslippers,whichwasobtainedbyaveragingthevaluesfromtheadjacentsimulationsymbols.Thefirststepinthedeterminationoftheverticalerrorbarsistocomputealowerandupperboundforthetransitionthreshold.Wedothisbydrawingalinethroughthehighestcroissantandlowestslippersymbols.ThisleadstothelowervioletanduppergreenlinesinfigureS5.Thus,thesetwolinesrepresenttheuncertaintyofthetransition,whichisaresultofthefinitedistancebetweenthesimulations.Anexceptionintheconstructionofthethreelinesoccursintheregionwherethesimulationspredictonlycroissants.Duetotheparticularstartingshape,thereisamaximalinitialoffset.Experimentally,however,itisofcoursepossiblethatthecellsatthechannelentrancehavealargeroffset(i.e.onethatliesabovetheblackdashedlineinfigureS5).Sincetheresultsfromthesimulationsindicatethatonlycroissantsreallyexistinthisregion(regardlessoftheinitialshapeandoffset),wetakertrans→∞.Thatwaywepredictavalueof1forthefractionofcroissants.Second,weneedtoevaluatethetransitionlinesattheexperimentalvelocities.However,themeasuredvelocitieshavenotonlyanaverageubutalsoacertainstandarddeviationσu.σuistakenastheuncertaintyinthevelocityhere.Evaluatingthemiddlegraylineattheaveragevelocityuresultsinthe"bestguessforrtrans"(thecircularsymbolsinfigureS5).Thisvalueisthendirectlyconvertedintothepredictedfractionofcroissantsφasdescribedinthemaintext(viaconversiontoytransandthemeasuredoffsetdistributionatthechannelentrance).Fortheverticalerrorbars,weevaluatethethreenumericaltransitionlines(lower,middleandupper,i.e.violet,grayandgreen)atthethreevelocitiesu,u−σuandu+σu,leadingtoninevaluesforrtrans.TheonesthatwillyieldthelowestandlargestfractionofcroissantsareshownastriangularsymbolsinfigureS5(the"lowestguessforrtrans"andthe"largestguessforrtrans").Third,thepredictedfractionsofcroissantsarecomputedfromtheoffsetdistributionatthechannelentranceforeachoftheseninertransvalues(asdescribedinthemaintext),andadditionallyforrtrans±sP.ThistakesintoaccounttheuncertaintyintheoffsetdistributionduetotheuncertaintysPinthepositionmeasurement.Asaresult,wenowhave27predictions.3LowestguessforrtransBestguessforrtransLargestguessforrtrans012345678910011.522.53∞Cellvelocity[mm/s]Initialradialpositionrinit[m]Initializedasdiscocytenon-TTCroissantTTCroissantnon-TTSlipperTTSlipperMax.offsetApproximatedtransitionMin.transitionMax.transitionApprox.trans.(atexp.velocity)Min.trans.(atexp.velocity)Max.trans.(atexp.velocity)FigureS5:Numericalphasediagramfromfigure6(a)fromthemaintextforthediscocytestartingshape.Thenearlytransparentshapesymbolsandthemaximaloffsetareidenticaltofigure6(a).Theviolet,grayandgreenlinesdepicttheminimal,averageandmaximalposition,respectively,ofthetransitionthresholdrtransbetweencroissantsandslippers.Theselinesareevaluatedattheexperimentalvelocitiesuandu±σu,givingthecircularandtriangularsymbols.Eachtripleofthesesymbolsthatsharesthesamecolorcorrespondstooneparticularexperimentalvelocityandshowsthelowest,bestandlargestguessforrtrans.Thisisexemplifiedviathethreelabelsandarrowsforu±σu=(3.16±0.14)mm/swhichcorrespondstoapressuredropof∆P=300mbar.Thehorizontalerrorbarsdepictσu.Alsonotethatinthecroissant-onlyregionwetakertrans→∞,asindicatedbythe∞symbolonthetopleft.4Fourth,wesearchfortheminimum(φmin)andmaximum(φmax)ofthese27values.φminandφmaxaretheninterpretedastheuncertaintyintheprediction.Theverticalerrorbarsinfigure8fromthemaintextthereforedepictφminandφmax.Allofthisisperformednotonlyforthephasediagramwiththediscocyte,butalsofortheoneswiththecroissantandslipperstartingshapes.Incaseofthecroissantstartingshape,rtransisnotaproperfunctionduetotheprotrusions,i.e.wefindseveraltransitionoffsetsforcertainvelocities(comparefigure6inthemaintext).Hence,thesimple"countingofcellsthatenterwithanoffsetbelowrtrans"toformthepredictionbecomesa"countingofcellsthatenterwithoffsetsintheintervalsformedbythenumericaltransitionoffsets".Asanexample,ifacertainvelocityleadstotransitionsatr1,r2andr3(suchthatthesimulationsyieldcroissantsinthetwointervals[0,r1]and[r2,r3]),thenwecounthowmanycellsenterthechannelwithanoffsetthatliesinthesetwointervals(aftertheirprojectiononthey-axis).Thecomputationoftheuncertaintyisadaptedaccordingly.S4AdditionalexperimentaldataWedepictinfigureS6themeasuredcellvelocitiesforeachappliedpressuredrop∆P.Thedatashowstheresultwhentheaveraginggoesoverallcellsregardlessoftheirshape("All"),andalsoforthethreeshapeclassesseparately.Obviously,thecellvelocitiesareroughlyproportionalto∆P.However,croissantstendtobeabitfasterthanslippersbecausecroissantsarelocatedinthehigh-velocitycenterofthechannelwhileslippersareoff-centered(seethemaintext).Thisisinagreementwithpreviouspublications[1,2].TableS1liststhecorrespondingrawdata,aswellasthenumberofcellsthatweretakenintoaccount.Itadditionallyshowsthenumberofcellsatthechannelentrance.TherawimagesfromtheexperimentsareincludedinsectionsS6andS7below.Furthermore,welistinanextraExcelsheet(SI_rawYPos_pos0.xls)therawy-positionsofthecellsatthechannelentrance.Thisdatamakesitpossibletocompareone'sownsimulationresultswithourexperiments(aswedidinfigure8inthemaintext).Moreover,figureS7depictstheexperimentaly-offsetdistributionsseparatedintothecontributionsfromthethreedifferentshapes(croissants,slippersand"others")atpositionx=10mminthechannel.Thisfigurecomplementsfigure3(b)fromthemaintextwhereallthreeshapeshavebeenconsideredtogether.02004006008001000024681012Appliedpressuredrop[mbar]Measuredcellvelocity[mm/s]AllCroissantsSlippersOthersFigureS6:Measuredaveragecellvelocitiesforeachappliedpressuredropforthethreedifferentshapeclassesandonceforallshapestogether("All").Theverticalerrorbarsdepictthestandarddeviationσu.Measurementsperformedatpositionx=10mminthechannel.ThecorrespondingrawdataislistedintableS1.Thelinesareguidesfortheeyes.S5References[1]S.Quint,A.F.Christ,A.Guckenberger,S.Himbert,L.Kaestner,S.Gekle,andC.Wagner,"3Dtomographyofcellsinmicro-channels,"Appl.Phys.Lett.111,103701(2017).[2]G.Tomaiuolo,M.Simeone,V.Martinelli,B.Rotoli,andS.Guido,"Redbloodcelldeformationinmicroconfinedflow,"SoftMatter5,3736(2009).5∆P[mbar]Nall0Nall10NCrois10NSlipper10NOther10uall10[mm/s]uCrois10[mm/s]uSlipper10[mm/s]uOther10[mm/s]203510790980.135±0.0210.132±0.0200.135±0.02150105290430.43±0.040.440±0.0050.43±0.04100292051652380.98±0.070.98±0.070.996±0.0020.99±0.0420071484252222102.07±0.102.09±0.082.03±0.062.06±0.12300954751021202533.16±0.143.19±0.203.10±0.093.18±0.1340090463801672164.19±0.194.33±0.144.08±0.154.23±0.1950017921517117815.2±0.45.46±0.165.16±0.115.2±0.76001511768124446.1±0.46.57±0.106.0±0.46.19±0.257001591230105187.3±0.77.3±0.77.3±0.7800752000169318.2±0.88.2±0.78.21±1.409001872822241399.3±1.210.17±0.069.3±1.39.6±0.3100014130502663910.6±0.910.5±1.010.7±0.6TableS1:Experimentaldata:Thetablelistsforeachappliedpressuredrop∆PthetotalnumberofanalyzedcellsNall0atpositionx=0mminthechannelandthetotalnumberofanalyzedcellsNall10atpositionx=10mm.Forthelatterwealsoshowthenumberofcroissants,slippersand"others",togetherwiththemeasuredvelocitiesu10.Theuncertaintiesarethestandarddeviation.Thesubscripts"0"and"10"intheheadingspecifythex-positioninthechannel(0mmor10mm).-2-1.5-1-0.5 0 0.5 1 1.5 22050100200300400500600Probability density [a.u.]y-offset from center [µm](a) Croissants-6-4-2 0 2 4 62003004005006007008009001000Probability density [a.u.]y-offset from center [µm](b) Slippers-6-4-2 0 2 4 620501002003004005006007008009001000Probability density [a.u.]y-offset from center [µm](c) OthersFigureS7:Estimatedprobabilitydensityfunctionsfortheexperimentaly-offsetdistributionsatpositionx=10mminthechannelfor(a)thecroissant,(b)theslipperand(c)the"other"shapes.Theresultforallthreeshapescombinedwasshowninfigure3(b)inthemaintext.Theareabeloweachcurveisnormalizedto1,andtheyareoffsetintheverticaldirectionforillustrationpurposes.Alsonotethedifferentscaleofthehorizontalaxisinthefirstfigure.6S6Rawexperimentalimagesatx=10mmNote:Theimagesintheindividualcollectionsareorderedfromcenteredtooff-centered.Note:Manyofthe"others"(e.g.for∆P=200mbar)mightbecroissants,buttheycanalsobeslippersthatareviewedfromthe"top"(i.e.whencamerawouldpointalongthez-direction,onemightseeslippers).Sincewecannotdecidethisfromtheseimages,weclassifythemas"others".∆P=20mbar,x=10mm∆P=50mbar,x=10mm∆P=100mbar,x=10mm7∆P=200mbar,x=10mm∆P=300mbar,x=10mm8∆P=400mbar,x=10mm∆P=500mbar,x=10mm∆P=600mbar,x=10mm9∆P=700mbar,x=10mm∆P=800mbar,x=10mm10∆P=900mbar,x=10mm∆P=1000mbar,x=10mm11S7Rawexperimentalimagesatx=0mmNote:Theimagesintheindividualcollectionsareorderedfromcenteredtooff-centered.∆P=20mbar,x=0mm∆P=50mbar,x=0mm∆P=100mbar,x=0mm∆P=200mbar,x=0mm∆P=300mbar,x=0mm∆P=400mbar,x=0mm∆P=500mbar,x=0mm12∆P=600mbar,x=0mm∆P=700mbar,x=0mm∆P=800mbar,x=0mm∆P=900mbar,x=0mm∆P=1000mbar,x=0mm13 |
1108.1924 | 1 | 1108 | 2011-08-09T13:35:57 | Phenomenological analysis of ATP dependence of motor protein | [
"physics.bio-ph",
"math-ph",
"math-ph",
"q-bio.SC"
] | In this study, through phenomenological comparison of the velocity-force data of processive motor proteins, including conventional kinesin, cytoplasmic dynein and myosin V, we found that, the ratio between motor velocities of two different ATP concentrations is almost invariant for any substall, superstall or negative external loads. Therefore, the velocity of motor can be well approximated by a Michaelis-Menten like formula $V=\atp k(F)L/(\atp +K_M)$, with $L$ the step size, and $k(F)$ the external load $F$ dependent rate of one mechanochemical cycle of motor motion in saturated ATP solution. The difference of Michaelis-Menten constant $K_M$ for substall, superstall and negative external load indicates, the ATP molecule affinity of motor head for these three cases are different, though the expression of $k(F)$ as a function of $F$ might be unchanged for any external load $F$. Verifications of this Michaelis-Menten like formula has also been done by fitting to the recent experimental data. | physics.bio-ph | physics | Phenomenological analysis of ATP dependence of motor protein
Yunxin Zhang∗
Laboratory of Mathematics for Nonlinear Science, Centre for Computational System Biology,
School of Mathematical Sciences, Fudan University, Shanghai 200433, China.
(Dated: November 8, 2018)
In this study, through phenomenological comparison of the velocity-force data of processive motor
proteins, including conventional kinesin, cytoplasmic dynein and myosin V, we found that, the ratio
between motor velocities of two different ATP concentrations is almost invariant for any substall,
superstall or negative external loads. Therefore, the velocity of motor can be well approximated
by a Michaelis-Menten like formula V = [ATP]k(F )L/([ATP] + KM ), with L the step size, and
k(F ) the external load F dependent rate of one mechanochemical cycle of motor motion in satu-
rated ATP solution. The difference of Michaelis-Menten constant KM for substall, superstall and
negative external load indicates, the ATP molecule affinity of motor head for these three cases are
different, though the expression of k(F ) as a function of F might be unchanged for any external load
F . Verifications of this Michaelis-Menten like formula has also been done by fitting to the recent
experimental data.
The processive motor proteins,
including kinesin,
dynein and myosin are essential for biophysical function-
ferent constant KM for substall, superstall and negative
external loads. So the limit velocity in saturated ATP
ing of eukaryotic cells [1, 2]. Due to the development
solution is V∗ = k(F )L, and the velocity corresponding
of experimental instrument [3, 4], much accurate exper-
to ATP concentration [ATP] can be obtained simply by
imental data have been obtained [4 -- 13]. Both conven-
multiplying V∗ by a constant [ATP] /([ATP] +KM ).
tional kinesin and cytoplasmic dynein move hand-over-
hand along microtubules with step size 8.2 nm by con-
verting chemical energy stored in ATP molecules into me-
chanical works [10, 14 -- 17]. Myosin (V or VI) also moves
hand-over-hand but along actin filament with step size
around 36 nm [8, 18 -- 20]. So far, there are many bio-
physical models to understand the mechanism of motor
proteins, including the flashing ratchet model [11, 21, 22],
Fokker-Planck equation [23 -- 25]. Meanwhile, more de-
tailed mechanochemical models have also been designed
to explain the experimental data, and get meaningful bio-
chemical parameters [13, 26 -- 31].
In this study, by phenomenological comparison of the
velocity-force data of different ATP concentrations, we
found that the velocity of processive motor proteins
can be described by a Michaelis-Menten like formula
V = [ATP]k(F )L/([ATP] + KM ), but might with dif-
For the sake of comparison, the velocity-force data of
kinesin, dynein and myosin are plotted in Figs. 1, 2
and 3(a). In Fig. 1(a), the thick dashed line V1 is the
velocity-force data of kinesin for [ATP] =1 mM obtained
by Nishiyama et al [6], and the solid line V2 is for [ATP]
=10 µM. One can easily see that there is only little differ-
ence between the lines V2 and V1/3.3. Similar phenomena
can also be found for the velocity-force data of dynein
and myosin obtained in [8, 10, 32], see Figs. 1(b,c,d).
Generally, these phenomena also exits for negative and
superstall force cases, but might with different ratio con-
stant, see Figs. 2 and 3(a) for data of kinesin obtained
in Refs.
[4, 7, 9]. For the kinesin data in [9], the ratio
constant is about 2.6 for F < 0, about 5.6 for 0 ≤ F ≤ 7
pN, and about 2.3 for F > 7 pN [see Fig. 2(a)]. For
the data in [7], the ratio constant is about 29 for F < 0,
and about 16 for 0 ≤ F ≤ 5 pN [see Fig. 2(b)]. But for
the kinesin data measured in [4], the constant 3.6 works
well for both substall and negative external load [see Fig.
∗Email: [email protected]
3(a)].
1
1
0
2
g
u
A
9
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
4
2
9
1
.
8
0
1
1
:
v
i
X
r
a
Data for kinesin from Nishiyama 2002
Data for dynein from Toba 2006
2
)
m
n
(
y
t
i
c
o
e
V
l
1000
800
600
400
1mM (V1)
V1/3.3
10µ M (V2)
200
(a)
0
0
2
4
F (pN)
6
Data for myosin from Uemura 2004
800
600
400
200
)
m
n
(
y
t
i
c
o
e
V
l
(b)
0
0
1mM (V1)
V1/6.5
10µ M (V2)
2
4
F (pN)
6
Data for Myosin from Tsygankov 2007
)
m
n
(
y
t
i
c
o
e
V
l
600
400
200
(d)
0
0
2mM (V1)
V1/13
1µ M (V2)
1
F (pN)
2
3
)
m
n
(
y
t
i
c
o
e
V
l
600
400
200
(c)
0
0
1mM (V1)
V1/4.5
10µ M (V2)
1
F (pN)
2
3
FIG. 1: For the positive substall external load cases, the ve-
locity V2 of motor protein at low ATP concentration can be
well approximated by the velocity V1 at high ATP concentra-
tion divided by a constant. (a) For the experimental data of
kinesin measured in [6], the velocity V2 of [ATP] =10 µM can
be approximated by V1/3.3 with V1 the velocity of [ATP] =1
mM. (b) For the data of dynein measured in [10], velocity V2
FIG. 2: For general external load cases, the velocity V2 of
kinesin at low ATP concentration can be well approximated
by the velocity V1 at high ATP concentration divided by a
constant. (a) For the data in [9], the velocity V2 of [ATP]
=10 µM can be well approximated by velocity V1 for [ATP] =1
mM divided by a constant Const with Const=2.6 for F < 0,
Const=5.6 for 0 ≤ F ≤ 7 pN, and Const=2.3 for F > 7
pN. (b) For the data in [7], the velocity V2 of [ATP] =4.2
µM can be well approximated by velocity V1 for [ATP] =1.6
mM divided by a constant Const with Const=29 for F < 0,
Const=16 for F ≥ 0.
dependent expression of rate k(F ) should be given firstly.
of [ATP] =10 µM can be well approximated by V1/6.5 with
Usually, the mechanical coupled cycle of ATP hydrolysis
V1 the velocity of [ATP] =1 mM. (c) For the data of myosin
includes several internal states, here, as demonstrated
V measured in [8], velocity V2 of [ATP] =10 µM can be well
in the previous mechanochemical model [27], we assume
approximated by V1/4.5 with V1 the velocity of [ATP] =1
that, in each cycle, there are two internal states, denoted
mM. (d) For the data of myosin V used in [32] (derived from
[5]), velocity V2 of [ATP] =1 µM can be well approximated
by V1/13 with V1 the velocity of [ATP] =2 mM.
From the above observations about the experimen-
tal data plotted in Figs.
1 and 2, one can see that
the velocity-force relation of motor proteins satisfies
V (F, [ATP]) = f ([ATP])V∗(F ). Where V∗ = V∗(F )
is the velocity-force relation at saturated ATP concen-
tration, and obviously V∗ can be written as V∗(F ) =
k(F )L with L the step size of motor protein, and k(F )
the force dependent rate to complete one ATP hydrol-
ysis cycle (coupled with one mechanical cycle). The
function f ([ATP]) increases with [ATP], f (0) = 0 and
f ([ATP]) = 1 with [ATP] → ∞. A reasonable form of
f ([ATP]) is f ([ATP]) = [ATP]/([ATP] + KM ) with a
parameter KM which we called Michaelis-Menten con-
stant. Finally, the velocity formula can be written as
V (F, [ATP]) = [ATP]k(F )L/([ATP] + KM ).
To verify the above velocity-force formula, the force
by state 1 and state 2 respectively.
one cycle
· · · ⇄
}
z
1 ⇄ 2 ⇄ 1 ⇄ · · ·
{
(1)
Let ui, wi be the forward and backward transition rates
at state i, then the steady state rate k(F ) can be obtained
as follows [27, 33]
k =
u1u2 − w1w2
u1 + u2 + w1 + w2
.
(2)
The force dependence of rates ui, wi are assumed to be
[27]
u1(F ) = u0
w1(F ) = w0
1e−θ+
1eθ−
1 F L/kB T ,
1 F L/kB T ,
u2(F ) = u0
w2(F ) = w0
2e−θ+
2eθ−
2 F L/kB T ,
2 F L/kB T .
(3)
Where kB is the Boltzmann constant, T is the absolute
temperature, and θ±
i are load distribution factors which
1 + θ−
0 + θ−
satisfy θ+
1 = 1. For this two-state model,
one can easily get the following formula of motor velocity
0 + θ+
V (F, [ATP]) =
[ATP](u1u2 − w1w2)L
([ATP] + KM )(u1 + u2 + w1 + w2)
.
(4)
Data for kinesin from Guydosh 2009
(a)
800
600
400
200
)
m
n
(
y
t
i
c
o
e
V
l
)
1
2mM (V
V
/3.6
1
10µM (V
)
2
(b)
900
700
500
300
100
0
)
m
n
(
y
t
i
c
o
e
V
l
TABLE I: Parameter values used in the theoretical predic-
tions of the velocity-force relation for conventional kinesin,
3
cytoplasmic dynein and myosin V: see Figs. 3(b) and 4(a)(b).
The unit of rate u0
i is s−1.
i , w0
0
−2
−1
0
F (pN)
1
2
3
−100
−15
−10
−5
0
F (pN)
5
10
15
u0
1
u0
2
w0
1 w0
2
θ+
1
θ−
1
θ+
2
θ−
2
FIG. 3: (a) For the kinesin data measured in [4], the veloc-
ity V2 of kinesin at low ATP concentration (10 µM ) can be
well approximated by the velocity V1 at high ATP concentra-
tion (2 mM) divided by a constant 3.6, which is the same for
both substall and negative external load.
(b)Experimental
data for conventional kinesin measured in [9] and the the-
4235.5
kinesin 716.6
dynein 910.8 1.15 × 104 64.0
myosin 584.0 1.73 × 104 2.55
0.25 13.5 -0.014 0.609 0.378 0.027
0
0
-0.019 0.019 0.386 0.614
0.03
0.43 0.03 0.51
Michaelis-Menten constant KM = 60.3 µM and 14.8 µM)
I for the corresponding pa-
respectively, see also Tab.
oretical prediction using the Michaelis-Menten like formula
rameter values [34]. Note, the step size used in the cal-
V = [ATP]k(F )L/([ATP] + KM ). The ATP concentrations
are corresponding to [ATP] =1 mM (dashed line and squares)
and 10 µM (solid line and dots) respectively. The model pa-
rameter KM is 15.8 µM for F < 0, 39.2 µM for 0 ≤ F ≤ 7
pN, and 11.9 µM for F > 7 pN, others are listed in Tab. I.
culations is L = 8.2 nm for motor proteins kinesin and
dynein, but L = 36 nm for myosin V. Certainly, the same
fitting process can also be done to other experimental
data. The plots in Figs. 3(b) and 4 indicate that, the
experimental data of motor proteins can be well repro-
800
600
400
200
)
m
n
(
y
t
i
c
o
e
V
l
(a)
600
500
400
300
200
100
)
m
n
(
y
t
i
c
o
e
V
l
0
0
1
2
3
4
F (pN)
5
6
7
8
0
0
0.5
1
1.5
F (pN)
2
2.5
3
(b)
duced by the Michaelis-Menten like formula (4), so our
phenomenological analysis about the ATP dependence of
motor motion is reasonable.
In summary, in this study, the ATP dependence of
motor proteins is phenomenological discussed. Based
on the recent experimental data and basic numerical
FIG. 4: Experimental data for cytoplasmic dynein ob-
calculations, we found the motor velocity can be well
tained in [10] and myosin obtained in [5] (see also [32] for
described by a Michaelis-Menten like formula V =
the method to get the present values), and the theoreti-
cal prediction using the Michaelis-Menten like formula V =
[ATP]k(F )L/([ATP] + KM ). (a) The experimental data are
for [ATP] =1 mM (dashed line and squares) and 10 µM (solid
line and dots). The model parameter KM = 60.3 µM for
0 ≤ F ≤ 7 pN. (b) The experimental data are for [ATP] =2
mM (dashed line and squares) and 1 µM (solid line and dots).
The model parameter KM = 14.7 µM for 0 ≤ F ≤ 3 pN.
The fitting results of the above velocity-force formula
to kinesin data measured in [9] are plotted in Fig. 3(b).
In which, the Michaelis-Menten constant KM = 15.8 µM
for F < 0, KM = 39.2 µM for 0 ≤ F ≤ 7 pN, and
KM = 11.9 µM for F > 7 pN, other parameter values
I. Meanwhile, the fitting results to
are listed in Tab.
[ATP]k(F )L/([ATP] + KM ) with force dependent rate
k(F ) at saturated ATP. The different values of KM for
load imply,
substall, superstall and negative external
the ATP molecule affinity of motor head might be dif-
ferent for these three cases, but the basic mechanism
in each mechanochemical cycle (either forward or back-
ward) might be the same. An obvious conclusion from
our Michaelis-Menten like formula is that the stall force,
under which the mean motor velocity is vanished, is in-
dependent of ATP concentration [9, 10, 12].
Acknowledgments
the dynein data measured in [10] and myosin data mea-
This study is funded by the Natural Science Founda-
sured in [5] are plotted in Fig. 4(a) and Fig. 4(b) (with
tion of Shanghai (under Grant No. 11ZR1403700).
4
[1] J. Howard. Mechanics of Motor Proteins and the Cy-
yses one ATP per 8-nm step. Nature, 388:386 -- 390, 1997.
toskeleton.
Sinauer Associates and Sunderland, MA,
[15] W. Hua, Edgar C. Young, Margaret L. Fleming, and Jeff
2001.
Gelles. Coupling of kinesin steps to ATP hydrolysis. Na-
[2] R. D. Vale. The molecular motor toolbox for intracellular
ture, 388:390 -- 393, 1997.
transport. Cell, 112:467 -- 480, 2003.
[16] Charles L. Asbury, Adrian N. Fehr, and Steven M. Block.
[3] W. J. Greenleaf, M. T. Woodside, and S. M. Block. High-
Kinesin moves by an asymmetric hand-over-hand mech-
resolution, single-molecule measurements of biomolecular
anism. Science, 302:2130 -- 2134, 2003.
motion. Annu. Rev. Biophys. Biomol. Struct., 36:171,
[17] Ahmet Yildiz, Michio Tomishige, Ronald D. Vale, and
2007.
Paul R. Selvin. Kinesin walks hand-over-hand. Science,
[4] N. R. Guydosh and S. M. Block. Direct observation of
303:676 -- 678, 2004.
the binding state of the kinesin head to the microtubule.
[18] A. Yildiz, J. N. Forkey, S. A. McKinney, T. Ha, Y. E.
Nature, 461:125 -- 128, 2009.
Goldman, and P. R. Selvin. Myosin-V walks hand-over-
[5] A. D. Mehta, R. S. Rock, M. Rief, J. A. Spudich, M. S.
hand: Single fluorophore imaging with 1.5-nm localiza-
Mooseker, and R. E. Cheney. Myosin-V is a processive
tion. Science, 300:2016 -- 2065, 2003.
actin-based motor. Nature, 400:590 -- 593, 1999.
[19] G. E. Snyder, T. Sakamoto, J. A. Hammer, J. R. Sell-
[6] Masayoshi Nishiyama, Hideo Higuchi,
and Toshio
ers, and P. R. Selvin. Nanometer localization of single
Yanagida. Chemomechanical coupling of the forward and
green fluorescent proteins: Evidence that Myosin V walks
backward steps of single kinesin molecules. Nat. Cell
hand-over-hand via telemark configuration. Biophys. J.,
Biol., 4:790 -- 797, 2002.
87:1776 -- 1783, 2004.
[7] Steven M. Block, Charles L. Asbury, Joshua W. Shaevitz,
[20] M. Iwaki, A. H Iwane, T. Shimokawa, R. Cooke, and
and Matthew J. Lang. Probing the kinesin reaction cycle
T. Yanagida. Brownian search-and-catch mechanism for
with a 2D optical force clamp. Proc. Natl. Acad. Sci.
myosin-VI steps. Nat. Chem. Biol., 5:403 -- 405, 2009.
USA, 100:2351 -- 2356, 2003.
[21] R. D. Astumian. Thermodynamics and kinetics of a
[8] S. Uemura, H. Higuchi, A. O. Olivares, E.e M. De La
brownian motor. Science, 276:917 -- 922, 1997.
Cruz, and S. Ishiwata. Mechanochemical coupling of two
[22] J. M. R. Parrondo, J. M. Blanco, F. J. Cao, and R. Brito.
substeps in a single myosin V motor. Nat. Struct. Mol.
Efficiency of brownian motors. Europhys. Lett., 43:248 --
Biol., 11:877 -- 883, 2004.
254, 1998.
[9] N. J. Carter and R. A. Cross. Mechanics of the kinesin
[23] H. Risken. The Fokker-Planck Equation.
Springer,
step. Nature, 435:308 -- 312, 2005.
Berlin, 1989.
[10] S. Toba, T. M. Watanabe, L. Yamaguchi-Okimoto, Y. Y.
[24] R. Lipowsky. Universal aspects of the chemomechanical
Toyoshima, and H. Higuchi. Overlapping hand-over-hand
coupling for molecular motors. Phys. Rev. Lett., 85:4401 --
mechanism of single molecular motility of cytoplasmic
4404, 2000.
dynein. Proc. Natl. Acad. Sci. USA, 103:5741 -- 5745, 2006.
[25] Y. Zhang. The efficiency of molecular motors. J. Stat.
[11] J. Christof M. Gebhardt, A. E.-M. Clemen, J. Jaud, and
Phys., 134:669 -- 679, 2009.
M. Rief. Myosin-V is a mechanical ratchet. Proc. Natl.
[26] M. Rief, R. S. Rock, A. D. Mehta, M. S. Mooseker, R. E.
Acad. Sci. USA, 103:8680 -- 8685, 2006.
Cheney, and J. A. Spudich. Myosin-V stepping kinetics:
[12] Arne Gennerich, Andrew P. Carter, Samara L. Reck-
A molecular model for processivity. Proc. Natl. Acad.
Peterson, and Ronald D. Vale. Force-induced bidirec-
Sci. USA, 97:9482 -- 9486, 2000.
tional stepping of cytoplasmic dynein. Cell, 131:952 -- 965,
[27] M. E. Fisher and A. B. Kolomeisky.
Simple
2007.
mechanochemistry describes the dynamics of kinesin
[13] Peiying Chuan, James A. Spudich, and Alexander R.
molecules. Proc. Natl. Acad. Sci. USA, 98:7748 -- 7753,
Dunn. Robust mechanosensing and tension generation
2001.
by myosin VI. J. Mol. Biol., 405:105 -- 112, 2011.
[28] A. B. Kolomeisky and M. E. Fisher. A simple kinetic
[14] Mark J. Schnitzer and Steven M. Block. Kinesin hydrol-
model describes the processivity of myosin-v. Biophys.
5
J., 84:1642 -- 1650, 2003.
[33] B. Derrida. Velocity and diffusion constant of a periodic
[29] S. S. Rosenfeld and H. L. Sweeney. A model of myosin v
one-dimensional hopping model. J. Stat. Phys., 31:433 --
processivity. J. Biol. Chem., 279:40100 -- 40111, 2004.
450, 1983.
[30] Y. Q. Gao. A simple theoretical model explains dyneins
[34] The value of KM obtained in Figs. 3(b) and 4 might not
response to load. Biophys. J., 90:811 -- 821, 2006.
be consistent with the ratio constant used in Figs. 1 and
[31] Y. Zhang. Three phase model of the processive motor
2, since the plots in Figs. 1 and 2 are just phenomenolog-
protein kinesin. Biophys. Chem., 136:19 -- 22, 2008.
ical illustration, and the ratio constants are obtained by
[32] D. Tsygankov and M. E. Fisher. Mechanoenzymes un-
rough estimation. For example, for the dynein data plot-
der superstall and large assisting loads reveal structural
ted in Fig. 4(a), KM = 60.3 µM means the ratio constant
features. Proc. Natl. Acad. Sci. USA, 104:19321 -- 19326,
between V1 and V2 is 6.6, but 6.5 is used in Fig. 1(b).
2007.
|
1707.05453 | 1 | 1707 | 2017-07-18T03:45:22 | Free Energy Calculations of Membrane Permeation: Challenges due to Strong Headgroup-Solute Interactions | [
"physics.bio-ph"
] | Understanding how different classes of molecules move across biological membranes is a prerequisite to predicting a solute's permeation rate, which is a critical factor in the fields of drug design and pharmacology. We use biased Molecular Dynamics computer simulations to calculate and compare the free energy profiles of translocation of several small molecules across 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) lipid bilayers as a first step towards determining the most efficient method for free energy calculations. We study the translocation of arginine, a sodium ion, alanine, and a single water molecule using the Metadynamics, Umbrella Sampling, and Replica Exchange Umbrella Sampling techniques. Within the fixed lengths of our simulations, we find that all methods produce similar results for charge-neutral permeants, but not for polar or positively charged molecules. We identify the long relaxation timescale of electrostatic interactions between lipid headgroups and the solute to be the principal cause of this difference, and show that this slow process can lead to an erroneous dependence of computed free energy profiles on the initial system configuration. We demonstrate the use of committor analysis to validate the proper sampling of the presumed transition state, which in our simulations is achieved only in replica exchange calculations. Based on these results we provide some useful guidance to perform and evaluate free energy calculations of membrane permeation. | physics.bio-ph | physics | Free Energy Calculations of Membrane Permeation:
Challenges due to Strong Headgroup–Solute Interactions
Department of Chemistry, University of Washington, Seattle, WA 98195
Nihit Pokhrel and Lutz Maibaum
Understanding how different classes of molecules move across biological membranes is a prerequi-
site to predicting a solute's permeation rate, which is a critical factor in the fields of drug design and
pharmacology. We use biased Molecular Dynamics computer simulations to calculate and compare
the free energy profiles of translocation of several small molecules across 1,2-dioleoyl-sn-glycero-3-
phosphocholine (DOPC) lipid bilayers as a first step towards determining the most efficient method
for free energy calculations. We study the translocation of arginine, a sodium ion, alanine, and a
single water molecule using the Metadynamics, Umbrella Sampling, and Replica Exchange Umbrella
Sampling techniques. Within the fixed lengths of our simulations, we find that all methods pro-
duce similar results for charge-neutral permeants, but not for polar or positively charged molecules.
We identify the long relaxation timescale of electrostatic interactions between lipid headgroups and
the solute to be the principal cause of this difference, and show that this slow process can lead
to an erroneous dependence of computed free energy profiles on the initial system configuration.
We demonstrate the use of committor analysis to validate the proper sampling of the presumed
transition state, which in our simulations is achieved only in replica exchange calculations. Based
on these results we provide some useful guidance to perform and evaluate free energy calculations
of membrane permeation.
I.
INTRODUCTION
Cell membranes are impermeable to many ions and
hydrophilic compounds. The tight packing of phospho-
lipids, sterols, and other membrane components together
with the hydrophobic interior of the membrane prevent
entry of foreign particles into the cell. There are, how-
ever, several important exceptions to this general rule,
including larger molecules such as cell penetrating pep-
tides (CPPs) that are capable of permeating the bilayer.
Understanding the mechanism and kinetics of membrane
permeation by these molecules facilitates the understand-
ing of pore formation [1, 2], passive translocation of so-
lutes [3, 4], and drug delivery across cell membranes [5].
As a consequence, there is an enormous interest in dis-
covering new biomolecules that can spontaneously diffuse
across lipid membranes.
One
characteristic
common to most membrane-
permeating solutes is that they are positively charged [6],
and CPPs fall into this category. They have gained much
attention due to their ability to deliver large molecular
cargoes into the interior of a cell [7]. In addition to CPPs,
peptides with positively charged amino acid residues like
arginine and lysine have also been studied in an attempt
to use them as drug carriers across lipid bilayers [8]. De-
spite the massive interest in these systems, the precise
interactions of these carrier peptides with the membrane
still remains unclear [9–11]. In order to design novel drug
carriers, a detailed understanding of how small solutes
such as individual amino acids move across lipid bilayers
is necessary. This will allow to predict which peptides can
act as vehicles for drug delivery across cell membranes.
Membrane partitioning is difficult to measure experi-
mentally because of the complexity of lipid bilayer sys-
tems [12–14]. Computational approaches such as Molec-
ular Dynamics (MD) simulations can provide atomistic-
level insights into interactions of amino acids with lipid
bilayers, complementing experimental studies [15, 16].
Free energy profiles are often calculated as a first step
towards understanding the permeation mechanism and
predicting the kinetics of translocation [17–23]. Con-
ventional MD is not well suited for free energy calcu-
lations of rare events that involve high energy barriers,
which include membrane translocation. These large bar-
riers restrict the simulation from efficiently sampling the
entire configuration space because the system remains
trapped in a local free energy minimum. A number of
accelerated sampling methods that bias the system to
sample otherwise inaccessible regions have been devel-
oped. These methods aid in correctly predicting the free
energy of many interesting biological processes. While
all these methods improve the statistical sampling and
reduce the amount of required computational resources,
it is often not apparent which method is best suited for a
specific system of interest. In the last five years consider-
able attention has been given to identifying the sources
of sampling errors, where the difficulty lies in evaluat-
ing the convergence of results and estimating the asso-
ciated uncertainty within these simulation techniques.
Neale and coworkers have extensively studied conver-
gence of free energy profiles of translocation of small so-
lutes across lipid bilayers, specifically arginine transloca-
tion across DOPC [24–26]. They identified hidden bar-
riers that depend on the interactions of the solute with
water and lipid molecules. Considering the growing inter-
est of quantifying uncertainty in free energy calculations,
we here compare the efficiency of three popular accelera-
tion methods in the context of passive membrane translo-
cation: well tempered metadynamics (WT-metaD) [27],
umbrella sampling (US) [28], and replica exchange um-
brella sampling (umbrella exchange, UE) [29]. Other
methods that are frequently used in this context include
7
1
0
2
l
u
J
8
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
3
5
4
5
0
.
7
0
7
1
:
v
i
X
r
a
Adaptive Biasing Force [30] and Thermodynamic Inte-
gration [31], but are not considered here. To compare
these three methods, we use them to compute the free
energy of translocation of arginine, a single sodium ion,
different forms of alanine, and water across a DOPC bi-
layer. It should be noted that the translocation mecha-
nism across a pure phospholipid bilayer might be different
from that across a biological membrane, where channel
proteins can facilitate the process [32]. However, study-
ing small solutes traversing across such model membranes
can provide valuable insights into general principles of
passive membrane translocation. In addition, the solute
molecules were chosen because they are well studied in
the literature [18, 33–38] and represent a class of com-
pounds that commonly appear in translocation problems
and span a range of sizes, shapes and hydrophobicities
[39]. An extensive list of translocation research can be
found in recent review papers [26, 40]. While most pre-
vious works focus on the mechanistic details of the per-
meation phenomenon itself, we concentrate on identify-
ing and diagnosing generic convergence issues.
In par-
ticular, we compute the committor distribution function
to check if our simulations accurately sample the transi-
tion state ensemble (TSE) of the translocation process.
We find that free energy calculations of positive and po-
lar solutes do not converge within typical timescales of
membrane simulations. The difficulty in relaxation of
electrostatic interactions between the solute and lipid
headgroups gives rise to hysteresis-like behavior, which
decays only after extensive UE equilibration. Based on
our results, we identify some useful diagnostic tools to
evaluate the accuracy of calculated free energy profiles of
membrane permeation.
II. METHODS
A. System Preparation
We performed all simulations using Gromacs 4.6.7 [41]
with the Plumed 2.1 plugin [42] under periodic boundary
conditions. Temperature and pressure were maintained
at 320 K and 1 atm using the Nose-Hoover thermostat
and Berendsen barostat, respectively. Long range elec-
trostatic interactions were computed using the fourth or-
der PME method [43] with a Fourier spacing of 0.12 nm.
The real space coulombic interaction was calculated up
to 1.0 nm. Van der Waals interactions were calculated
using a cutoff of 1.0 nm. Bond lengths within the solutes
and lipids were constrained using the LINCS algorithm
[44].
The 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC)
bilayer was constructed using the united-atom Berger
forcefield [45] such that each monolayer consisted of 64
lipids. Water molecules were treated using the rigid sim-
ple point charge (SPC) model [46]. Each bilayer-water
system was equilibrated for 10 ns before adding any per-
meant to the system. The permeants were modeled using
2
the all-atom OPLS-AA forcefield [47]. Cationic arginine
3 and COO−) was constructed
with charged termini (NH+
using the pdb2gmx tool of Gromacs. A system contain-
ing a single sodium ion was built using the genion tool
of Gromacs. Three different forms of alanine were con-
structed: the first by truncating the side chain at the β-
carbon with the α-carbon replaced by a hydrogen; we call
this form the side chain analog, where the alanine residue
essentially becomes a methane molecule. This method
of truncating amino acids has been used in the past to
study amino acid interactions with model bilayer systems
[17, 48]. The second form of alanine was constructed with
neutral termini (NH2 and COOH). We made the third
3 and COO−), a charge
form with charged termini (NH+
neutral but zwitterionic molecule. Thus we have studied
the following permeants:
i. Arginine
ii. Sodium ion
iii. Side chain analog of alanine
iv. Alanine with neutral termini
v. Zwitterionic alanine
vi. Water
For the simulations of positively charged solutes a sin-
gle chloride ion was added to achieve overall charge
neutrality. Each permeant/water system was equili-
brated for 10 ns. The equilibrated water/bilayer and wa-
ter/permeant systems were then combined to form the
final water/bilayer/permeant system, which was again
equilibrated for 10 ns before any production run un-
der NPT conditions. The Visual Molecular Dynamics
(VMD) software was used to monitor and visually in-
spect all simulation trajectories [49].
B. Well Tempered Metadynamics (WT-metaD)
Metadynamics is a biasing technique that overcomes
sampling problems by adding a history-dependent bias
potential VG(z, t) to the collective variable z, which itself
is a function of the positions of the atoms in the simula-
tion [50]. The bias potential is constructed of Gaussian-
shaped energy hills that are deposited along the collective
variable,
VG(z, t) = ω0
exp
,
(1)
(cid:88)
t(cid:48)<t
(cid:18)
− (z − z(t(cid:48)))2
2σ2
(cid:19)
where ω0 is the height of the added Gaussian, σ is its
width, and z(t(cid:48)) is the value of the collective variable at
time t(cid:48). Each hill is deposited at a predefined rate and
centered at a previously explored configurations, biasing
the system towards configurations that have not yet been
explored. At longer times, the sum of added hills can be
used to calculate the unbiased free energy profile of the
variable z. Well-tempered metadynamics (WT-metaD) is
an improved form of metadynamics where the height ω0
of the hills decreases in previously visited regions, which
guarantees correct convergence of the free energy profiles
[27, 51]. The choice of the metadynamics parameters is
crucial for successful convergence [52]. The values we use
for this work can be found in Table S1.
We choose the normal component of the distance vec-
tor between the center of mass of the solute and the
center of mass of the lipid bilayer as the collective vari-
able z. We used snapshots from a WT-metaD trajectory,
with permeants at various positions relative to the bi-
layer, as the initial configurations for both US and UE
calculations; the snapshots were taken when the solute
first reached the desired distance after a minimum of 100
ns of WT-metaD simulation time.
C. Umbrella Sampling (US)
Umbrella Sampling also adds a biasing potential to the
system's Hamiltonian to enhance the sampling of config-
urations that are high in free energy [28]. In this case
the biasing potential is static. We choose a sequence of
"windows" that span the range of interest of the collec-
tive variable z. In the ith window the system is biased
to remain close to a predetermined value zi by using a
harmonic umbrella potential
Vi(z) =
k(z − zi)2,
1
2
(2)
where k is the stiffness of the harmonic potential. The
results of N independent simulations, each performed
with a different value of zi, are then combined using the
weighted histogram analysis method (WHAM) to obtain
an estimate of the unbiased free energy profile F (z) [53].
Uncertainties were estimated using bootstrapping analy-
sis as implemented in the g wham tool of the Gromacs
simulation suite [54].
We use the same reaction coordinate z for the US
calculation as for WT-metaD. For each solute we used
a series of windows with spacing of 0.1 nm spanning
the entire bilayer. A harmonic potential of k = 1000
kJ/mol/nm2 was used. 40 windows were used for argi-
nine simulation and each window was run for 100 ns,
resulting in 4.0 µs of total simulation time. 50 windows
were constructed for all three forms of alanine. For the
side chain analog and alanine with neutral termini, each
window was simulated for 20 ns totaling 1 µs of simula-
tion time. Zwitterionic alanine was simulated for 2.75 µs
where each window was run for 55 ns (see table S2).
D. Replica Exchange Umbrella Sampling (UE)
UE is very similar to US, with the addition that neigh-
boring windows can exchange their configurations.
In
UE, N parallel and independent simulations of the same
system, biased at different values of zi, are run, each of
these N simulations is called a replica. An exchange of
3
configurations between neighboring replicas is attempted
at a pre-determined frequency, and is accepted or rejected
based on the Metropolis criterion. This technique is sim-
ilar to other Replica Exchange schemes where different
replicas are simulated at different temperatures, which
improves the sampling at low temperatures by incor-
porating enhanced sampling at higher temperatures[55].
Free energy profiles and the associated uncertainties were
again computed using the g wham tool.
To allow for accurate comparisons between the simu-
lation methods we use the exact same parameters for US
and UE. Each replica was run for 165 ns for arginine and
for 55 ns for alanine. For sodium and water, each replica
was simulated for 27 ns and 8 ns, respectively. An ex-
change was attempted every 2 ps for all UE calculations.
Table S2 contains information about simulation lengths
for each solute using each of the three methods.
Hydrophilic solutes are most likely found either in bulk
water or at the water/membrane interface, which corre-
sponds to a local minimum in the free energy profile F (z).
Because the membrane is symmetric in our simulations,
there are two degenerate minima corresponding to the
upper (A) and lower (B) half-spaces that are separated
by the membrane. For a given configuration we define
the committor pB as the probability that a trajectory
initiated from this configuration with random initial ve-
locities will reach the free energy minimum B on the lower
(z < 0) side of the membrane before it reaches minimum
A on the upper side (z > 0). The collection of configu-
rations with pB = 1/2 form the transition state surface
that separates states that are more likely to go to A from
those that are more likely to go to B. This definition of
the transition surface does not necessarily correspond to
a saddle point in a free energy landscape. It is preferred
because of its intuitive kinetic interpretation and because
it does not require a choice of a reaction coordinate in
the transition region [56, 57]; instead it requires only a
well-defined reactant and product state. In this picture
a transition state can also be a metastable intermediate.
Given the symmetric nature of our membrane, one
might expect configurations in which a hydrophilic per-
meant is at the center of the bilayer to be transition
states. To test this hypothesis we calculate the distri-
bution of committors, P (pB), over multiple states with
z = 0. If all configurations in this ensemble are indeed
transition states, then P (pB) will be sharply peaked at
pB = 1/2 [56].
To compute the committor distribution function P (pB)
for an ensemble of configurations with z = 0 as gener-
ated by the WT-metaD, US, and UE methods we take
six such configurations from each of the three methods,
and initiate four unbiased trajectories with random ini-
tial velocities from each of these eighteen configurations.
We then count in how many of these four trajectories the
solute reaches the lower (z < 0) membrane/water inter-
face before the upper interface. The resulting fraction
serves as an estimate for the pB-value of a configuration,
and we construct a histogram of these values over the
4
FIG. 1. Free Energy of arginine translocation as a function
of distance from the bilayer center using WT-metaD, US and
UE with shaded error bars for US and UE. All three profiles
have two minima: one at the interface of upper leaflet and
water (IA) and one at the interface of the lower leaflet and
water (IB). The three lines differ: profiles generated using
WT-metaD and US are asymmetric and the maximum does
not occur at z = 0. The UE profile, obtained after extensive
equilibration, is symmetric and has its maximum at z = 0.
six configurations obtained from each sampling method.
These histograms are shown in Fig. 3.
Both the number of configurations sampled from the
z = 0 ensemble and the number of trajectories initiated
from these configurations to estimate the committor are
quite small. One typically needs better sampling to ob-
tain accurate estimates of pB and P (pB) [57]. However,
the long timescale of solute motion across the membrane
limits us to such relatively small numbers.
III. RESULTS
A. Committor Distribution Function
We compare the free energy profiles of translocation
for arginine, a single sodium ion, alanine, and a water
molecule as calculated by the WT-metaD, US and UE
methods. When calculating such profiles for symmet-
ric membranes, it is tempting to take advantage of this
symmetry and to calculate the free energy only for one
half-space, either z ≥ 0 or z ≤ 0, and obtain the other
half by simply mirroring the result with respect to the
z = 0 axis. Another way to exploit this symmetry is to
compute the free energy profile for the whole range of
z, and then average together the free energies above and
below the membrane center. Both approaches are for-
mally correct and yield symmetric free energy profiles by
construction. However, we will see that they can conceal
signatures of poor convergence of the simulations. We
therefore choose to compute free energy profiles for the
entire range of the collective variable z. This prevents us
from losing any information regarding the translocation
process and also helps us evaluate the discrepancy, if any,
FIG. 2. UE results for arginine. (a) Exchange pattern lines
for 40 replicas. Replicas starting at z = 2.0 nm, 0 nm and -1.9
nm are highlighted. Initial configurations for these replicas
are shown in ((c)-(e)) with water shown in red, headgroup
phosphates in orange, membrane outline in dark yellow, and
the solute at various position across the DOPC bilayer. A gap
in the exchange pattern (indicated by the arrow) is clearly
visible. Using the final configuration of the z = 0 replica
(shown in (f)) as the initial condition for the center replicas
in a new UE simulation (b), this gap is no longer present,
indicating even exchanges between all 40 replicas.
between the free energy profiles calculated using a single
versus both monolayers, as suggested previously [23, 25].
B. Arginine
In Fig. 1 we show the free energy of arginine translo-
cation across a DOPC bilayer as a function of arginine
distance from the membrane center as computed by WT-
metaD, US, and UE. Each profile is shifted vertically to
have the global minimum at zero. We see that all three
profiles have local minima at the interface of the bilayer
and water (z ≈ ± 1.7 nm), which indicates that arginine
prefers the interface over both the hydrophobic core and
the bulk water. At these minima, arginine forms strong
electrostatic interactions with lipid headgroups and re-
mains solvated by water.
Other than the position of these local minima, there
are stark differences between the three profiles. This is
unexpected as all three methods in principle converge to
the same free energy function. Furthermore, WT-metaD
and US do not yield symmetric profiles, as one would
expect for symmetric bilayers such as the one used in
this study. For example, our WT-metaD calculations
050100150200-3-2-10123FreeEnergy[kJ/mol]Distancefrombilayercenter[nm]IAIBWT-metaDUSUEReplicacenter[nm]SimulationTime[ns](a)gapSimulationTime[ns](b)-10120100200300050100150(c)(d)(e)(f)5
free energy profile. The obtained uncertainties are on the
order of 15kJ/mol for our US simulations, which severely
underestimates the actual error of the calculation. This
is not surprising, as such error estimates can only use in-
formation from the sampled trajectories, and by nature
have no knowledge of hitherto unsampled configurations.
Visual inspection of the simulation centered at z = 0
nm shows that throughout the simulation arginine is in
contact with lipid headgroups from the upper leaflet only.
This persistent association with molecules from only one
leaflet is an artifact of the initial configuration caused by
insufficient equilibration. We constructed the initial con-
figuration by pulling arginine from bulk water into the
membrane using WT-metaD. Our observation suggests
that even after 4 µs (100ns/window) of US simulation,
the system still remembers its initial configuration and
is therefore not equilibrated. This result is also consis-
tent with previous studies of translocation of arginine
side chain analogs across DOPC bilayers where sampling
error persisted for 125 ns [24].
The UE simulation method provides additional infor-
mation, not present in WT-metaD or US calculations,
that aid in the detection of such convergence failures.
In principle each trajectory should diffuse through the
complete space of replicas due to the ongoing exchange
between neighboring replicas. Fig. 2(a) visualizes the tra-
jectories of 40 replicas, spanning the bilayer from −2.0
nm< z ≤ 2.0 nm. Because exchange occurs frequently
between most replicas, it is difficult to discern individual
trajectories, and we highlight three specific ones to illus-
trate the motion of trajectories through replica space.
There is, however, a well-defined and persistent gap in
the exchange pattern, which is visible as a white line in
Fig. 2(a). This gap indicates that the two neighboring
replicas do not exchange configurations over long periods
of simulation time. The location of this gap is shifting
slowly on the timescale of the simulation. Its effect on
the exchange process can be seen by the highlighted tra-
jectory that starts in the replica centered at z = 2 nm:
it diffuses freely though replica space, but when it hits
the gap at approximately 260 ns of simulation time, it
cannot cross to the other side of the gap; instead it is re-
flected back towards replicas centered at positive values
of z. This gap separates replica space into two regions
that do not exchange configurations with each other.
Visual inspection reveals that this gap separates con-
figurations based on the association of the arginine solute
with lipid headgroups from the two different leaflets. In
replicas above the gap, the solute is in close contact with
lipid headgroups from the upper leaflet (panels (c) and
(d) of Fig. 2), whereas in trajectories below the gap it is
in contact with either both or just the bottom leaflet
(Fig. 2(e)). At approximately 280 ns the gap moves
across the z = 0 replica. After this jump, arginine is
in close contact with lipid headgroups from both leaflets
(Fig. 2(f)). Only after a long simulation time does the
z = 0 replica lose the memory that its initial configura-
tion was chosen in a way that the solute had a higher
FIG. 3. Committor histogram for 3 ensembles with arginine
at z = 0 nm. The single peak at pB = 0.5 for UE shows that
arginine has no preference for either side of the membrane in
all six tested configurations. Peaks at pB = 0 and pB = 1 for
WT-metaD and US show that each of the tested configura-
tions has a strong bias towards one side of the membrane.
predict that the free energy is highest at z ≈ −0.7 nm,
and that the free energy difference to the upper mem-
brane/water interface is 170 kJ/mol while that to the
lower interface is only 50 kJ/mol. If we had calculated
the free energy profile only for the z ≥ 0 half-space and
symmetrized it, we would have predicted a free energy
maximum of 90 kJ/mol and would have lost the infor-
mation about the asymmetry. Such a profile would also
have a sharp kink at z = 0 nm, which would indicate a
discontinuity in the mean force which is physically un-
likely. This shows that mirroring the free energy pro-
file at z = 0 nm or symmetrizing it across the bilayer
can be dangerous short-cuts that can yield misleading
results. Visual inspection of the WT-metaD trajectory
shows that phosphate groups from lipid headgroups in-
teract strongly with the arginine and are pulled along as
the solute is driven across the bilayer by the metadynam-
ics algorithm. This displacement of lipid molecules due
to strong interactions is likely the origin for the observed
asymmetric free energy profile.
It has been previously
observed by Neale and co-workers who studied translo-
cation of n-Propylguanidium across a DOPC membrane.
During translocation of the solute, the membrane forms
a depression that facilitates the continued interaction of
the bilayer surface with the solute [24].
US results show similar characteristics to results ob-
tained with WT-metaD. The maximum of the barrier
does not occur at z = 0 nm and the free energy differ-
ences between the maximum and the upper and lower
interfaces are different. The barrier height, however, is
smaller than that predicted by WT-metaD. It is impor-
tant to note that even sophisticated error analysis tech-
niques such as bootstrapping cannot accurately quantify
the discrepancy between the computed and the actual
012345600.51Number of configurationspBWT-metaDUSUEaffinity towards the upper leaflet.
Given the asymmetric profiles generated using WT-
metaD and US, we hypothesized that the configuration
shown in Fig. 2(f) is more representative for the equi-
librium ensemble of the z = 0 replica. We therefore
performed a new UE calculation using this configura-
tion as the starting point for replicas inside the bilayer.
Fig. 2(b) shows the exchange patterns of the replicas
for this second UE calculation.
It no longer exhibits
an apparent gap such as the one visible in Fig. 2(a),
and replicas can explore the entire replica space. This
UE simulation yields the symmetric free energy profile
shown in Fig. 1, with a maximum of 50 kJ/mol at the
center of the bilayer (z = 0) as one would expect from
a converged free energy calculation.
In this case, con-
vergence is also indicated by the very small magnitude
of the estimated uncertainty. Such a symmetric profile
was also observed by Neale and coworkers, who calcu-
lated the free energy profile for translocation of an argi-
nine side chain analog across a 1-palmitoyl-2-oleoyl-sn-
glycero-3-phosphocholine (POPC) bilayer using Virtual
Replica Exchange [25]. They attributed the symmetry of
the free energy to better sampling allowed by the ability
of replicas to sort themselves along the order parameter.
The free energy we calculate is lower than previously re-
ported values of arginine translocation across DOPC bi-
layers [18, 33]. These studies considered the side chain
analog of arginine, while we study the entire amino acid
residue, which is the likely reason for this discrepancy.
All reported free energy barriers for the translocation of
a single arginine residue or its analogs are significantly
higher than those found in experimental [39, 58] and
simulation [34] studies of arginine insertion as part of
a trans-membrane peptide.
To further test whether the success of the second UE
calculation and the failure to converge of the WT-metaD
and US calculations are related to an erroneous bias of
the solute towards either side of the membrane we cal-
culate the committor distribution P (pB) of the z = 0
ensemble generated by these three methods. The com-
mittor pB of a configuration is the probability that in
a trajectory initiated from this configuration with ran-
domized velocities, the solute reaches the lower mem-
brane/water interface before it reaches the upper one.
By definition, configuration with pB = 0.5 are transition
states. Intuitively one might expect that many (if not all)
equilibrium configurations in which the solute is at the
center of the bilayer have a committor value of 0.5, i.e.,
the solute is equally likely to go to the upper or the lower
interface. In this case a histogram of pB values over an
ensemble of such configurations should be sharply peaked
at pB = 0.5.
Details of the committor calculation are given in the
Methods section, and results are shown in Figure 3. The
committor histogram for the UE ensemble is peaked at
0.5 but the distributions for WT-metaD and US are not.
The former indicates that the second UE calculation cre-
ates an unbiased ensemble of configurations that have
6
FIG. 4. Free Energy of sodium translocation as a function
of distance from the bilayer center using WT-metaD and UE
with shaded error bars for UE. The two lines differ: profiles
generated using WT-metad is asymmetric and the maxima
does not lie at z = 0 nm. UE profile after extensive equili-
bration is symmetric with maximum at z = 0 nm.
no preference towards either leaflet. Configurations with
z = 0 as sampled by the WT-metaD and US methods,
on the other hand, have a significant bias towards one
leaflet over the other. For the WT-metaD ensemble, 3
out of the 6 initial configurations we chose occurred while
the arginine was crossing the bilayer from top to bot-
tom, and all these configurations returned to the upper
membrane/water interface in all four trajectories. In the
remaining 3 initial configurations the solute crossed the
membrane from bottom to top, all of those returned to
the bottom leaflet when four trajectories were initiated
from them. Together, this gives rise to the double-peaked
histogram shown in Fig. 3. It shows that in a configu-
ration in which the solute is at the bilayer center en-
countered along a WT-metaD trajectory, the solute has
a strong tendency to return to the side from which it just
came, an effect similar to hysteresis. All configurations
sampled by the US algorithm, on the other hand, exhib-
ited a strong bias towards the upper membrane/water
interface, resulting in a histogram with a single peak at
pB = 0. The likely reason for this is that the initial con-
dition for the z = 0 window in the US calculation was
taken from the WT-metaD trajectory where the arginine
had been pulled into the membrane from above. The
large peak at pB = 0 shows that the system remembers
this initial configuration even after 100 ns of simulation
time.
Irrespective of the method used, capturing the transi-
tion state and generating the free energy profile for argi-
nine translocation across a DOPC bilayer is challenging
and requires significant simulation time. The UE method
has the advantage that it provides a useful diagnostic tool
to identify potential convergence issues: one can study
the pattern of replica exchanges to check for unexpected
behavior of the simulation, such as the gap shown in
Fig. 2(a). This is an important advantage of this method
over the others, even though it is not directly related to
020406080-2-1012FreeEnergy[kJ/mol]Distancefrombilayercenter[nm]WT-metaDUE7
FIG. 5. Free Energy of three forms of alanine translocation as a function of distance from the bilayer center with shaded error
bars for US and UE. WT-metaD and US profiles are comparable for side chain analog of alanine (a) and alanine with neutral
termini (b). For zwitterionic alanine (c), the three lines differ: profile generated using WT-metad and US are asymmetric and
the maxima does not lie at z = 0 nm. UE profile after extensive equilibration is symmetric with maximum at z = 0 nm.
its sampling efficiency.
C. Sodium Ion
To test whether strong electrostatic interactions be-
tween zwitterionic lipid head groups and a charged so-
lute are the reason for long relaxation timescales we
performed WT-metaD and UE simulations of a single
sodium ion translocating across the membrane. The re-
sulting free energy profiles, shown in Figure 4, resemble
those of the arginine calculations: that obtained from
WT-metaD is not symmetric across the bilayer, and has
a maximum that is slightly offset from the bilayer center.
Inspection of our WT-metaD simulation trajectory re-
veals that lipid molecules whose headgroups are in close
contact with the ion are pulled along as the latter is
driven across the membrane. US can yield similar, asym-
metric free energy profiles for sodium translocation, as
reported in the literature [35].
The replica exchange pattern of the UE simulation
shows a persistent gap that slowly moves through replica
space (Fig. S1(a)), similar to that seen in the arginine
calculations. This gap separates replicas based on which
phospholipid headgroups are in close contact with the
sodium ion. For example, in the initial condition of the
z = −0.2 nm replica the sodium ion was surrounded
by headgroups from upper leaflet lipids. After approx-
imately 35 ns the gap crosses this replica, which from
that point onward contains conformations in which head-
groups from both leaflets are in contact with the ion. As
before, starting a new UE calculation from those config-
urations generates an exchange pattern without a persis-
tent gap (Fig. S1(b)), and yields a symmetric free energy
profile (Fig. 4). We find that the change in free energy of
moving the ion from the membrane/water interface to the
center of the bilayer is approximately 50 kJ/mol, which is
surprisingly close to the value we obtained for arginine. A
similar observation was made by Vorobyov and coworkers
who found that translocation free energy barriers are sim-
ilar for an arginine side chain analog and simple ions in
1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) bi-
layers [59].
D. Alanine
Both arginine and sodium are cationic. To test
whether the presence of a net charge on the solute is a
prerequisite for the long equilibration timescales that we
observe, we study the translocation free energy of three
different forms of alanine. The first is its side chain ana-
log, which is constructed by replacing the peptide back-
bone including the α-carbon by a single hydrogen atom;
this form is identical to a methane molecule. The second
form of alanine is built using charge-neutral peptide ter-
mini NH2 and COOH. The third form has the charged
3 and COO−. All three forms have no net
termini NH+
charge, but differ in the dipole moments of 0, 4.02 D,
and 14.4 D, respectively, as calculated by the g dipoles
tool of Gromacs.
The side chain analog of alanine has neither a net
charge nor a dipole moment, and therefore weak elec-
trostatic interactions with phospholipid headgroups. As
such one might expect that it does not stay in close
contact with them as the solute is driven through the
membrane, and that all methods should give comparable
results even without extensive equilibration. Fig. 5(a)
shows that this is indeed the case: both WT-MetaD and
US yield approximately symmetric free energy profiles
that are comparable to each other within error bars. One
key difference, which is unrelated to the used simulation
method, is that the side chain analog is hydrophobic, and
the free energy therefore has a minimum at the center of
0102030-2-1012FreeEnergy[kJ/mol]Distancefrombilayercenter[nm]SideChainAnalog(CH4)(a)0102030-2-1012Distancefrombilayercenter[nm]Alaninewithneutraltermini(NH2CHCH3COOH)(b)0306090120-2-1012Distancefrombilayercenter[nm]ZwitterionicAlanine(NH+3CHCH3COO−)(c)WT-metaDUSWT-metaDUSWT-metaDUSUEthe bilayer.
For the second form of alanine with neutral termini,
its dipole moment is sufficiently large to prefer the bulk
water phase or the membrane/water interface region over
the membrane interior, but not large enough to induce
strong interactions with lipid headgroups. Therefore all
methods yield converged results, as shown in Fig. 5(b)
for WT-MetaD and US.
Sampling the free energy profile of zwitterionic ala-
nine, on the other hand, reveals similar behavior previ-
ously observed for the charged arginine and sodium so-
lutes. Strong interactions between the solute and lipid
headgroups cause the latter to be dragged along as the
alanine is driven towards the interior of the bilayer by the
WT-MetaD algorithm. These interactions do not relax
on the simulation timescale, which leads to the asymmet-
ric shapes of non-converged free energy profiles for WT-
MetaD and US (Fig. 5(c)). As was the case in the argi-
nine calculation, the estimated error of the US profile is
far too small, which is most likely due to insufficient sam-
pling throughout the simulations. The existence of the
electrostatic bottleneck can be detected by studying the
exchange pattern of UE calculations, which exhibit a per-
sistent gap indicating non-overlapping neighboring repli-
cas (Fig. S2(a)). Visual inspection of the trajectories
reveals that replicas inside the bilayer initially contain
configurations in which the solute is in close contact with
lipid headgroups from only one leaflet, but transitions to
configurations in which both leaflets' headgroups are in
contact with the solute once the gap has moved passed
the replica. Using the latter conformations as the initial
condition for a new UE calculation yields exchange pat-
terns without persistent gaps (Fig. S2(b)) and a nearly
symmetric, well-converged free energy profile (Fig. 5(c)).
We calculate a barrier height of 60 kJ/mol at the center
of the bilayer.
These results show that long relaxation timescales and
the resulting sampling problems are not unique to the
translocation of charged solutes. Permeants with a suffi-
ciently large dipole moment can exhibit the same behav-
ior, and the analysis of UE exchange patterns can aid in
the detection and resolution of such problems.
E. Water
Finally we calculate the free energy profile of translo-
cation of a single water molecule across the bilayer.
Within the SPC model used in our simulations a water
molecule has a dipole moment of 2.27 D, which based on
our previous analysis of alanine we expect is sufficiently
small to avoid sampling difficulties due to long relaxation
timescales. This is verified by the lack of a persistent gap
in the UE exchange pattern (Fig. S3) and the convergence
of both WT-metaD and UE calculations to symmetric
free energy profiles within error bars(Fig. 6). We find a
barrier height of 27 kJ/mol at the bilayer center, which is
in good agreement with previous studies of water translo-
8
FIG. 6. Free Energy of a single water molecule as a function
of distance from the bilayer center using WT-metaD and UE.
Uncertainties were computed using bootstrapping for UE re-
sults and are indicated by the shaded area. The two results
are comparable, in particular both free energy profiles are
symmetric with respect to the bilayer center.
cation through PC bilayers using the US method [36, 37].
Visual inspection of the WT-metaD trajectory confirms
that no lipid headgroups are dragged along the water
molecule as it is driven through the membrane.
IV. DISCUSSION
Our results demonstrate that the calculation of accu-
rate translocation free energy profiles of solute motion
across lipid bilayers can require long simulation times and
care in the choice of initial conditions for all tested simu-
lation methods. This is the case for the positively charged
solutes arginine and sodium as well as the strongly dipo-
lar zwitterionic alanine.
In these cases strong electro-
static interactions between the solute and phospholipid
headgroups create a long-lived affinity of the solute to
one bilayer leaflet even when the solute is in the cen-
ter of the bilayer that takes hundreds of nanoseconds to
relax. Similar timescales related to membrane reorga-
nization have been observed in simulations of peptides
interacting with membranes [60–62]. Small, uncharged
solutes with no (side chain analog of alanine) or small
(neutral alanine, water) dipole moments do not exhibit
this dynamical bottleneck.
Neither of the three tested simulation methods can in-
trinsically accelerate this relaxation process. However,
replica exchange umbrella sampling (UE) has a desirable
feature lacking both in umbrella sampling (US) and well-
tempered metadynamics (WT-metaD): by analyzing the
exchange pattern between replicas one can easily detect
sampling deficiencies caused by mismatched configura-
tions in neighboring replicas. These manifest themselves
as persistent gaps in the exchange pattern that hinders
efficient mixing in replica space. Movement or disappear-
ance of this gap can aid in choosing initial conditions for
subsequent simulations that do not suffer from this bot-
0102030-3-2-10123FreeEnergy[kJ/mol]Distancefrombilayercenter[nm]WT-metaDUEtleneck. The importance of even exchanges between all
replicas has been previously emphasized by Neale and
coworkers, who developed a novel metric, the transmis-
sion factor, to quantify the diffusivity of replicas [25].
In all cases it is advisable to compute the free energy
profile for translocation across the entire lipid bilayer.
Calculating the profile only for one half-space and then
mirroring it to obtain the other is sufficient in princi-
ple and minimizes computing cost, but in practice leads
to the loss of valuable information: for symmetric mem-
branes, an asymmetric free energy profile is an indicator
that the calculations are not yet converged. Similarly,
symmetrizing a free energy profile obtained for the en-
tire membrane should not be done blindly, as it might
lead to an unwarranted confidence in the result.
In this work he have focused only on a single collective
variable: the normal distance between the bilayer center
and the solute. This is by far the most widely used order
parameter in free energy calculations of solute-bilayer in-
teractions. Choosing different (or additional) collective
variables for the biased sampling schemes might allevi-
ate the convergence issues we encountered. However, the
selection of the optimal order parameter in itself is a chal-
lenging problem and beyond the scope of this paper [63–
65].
Another field of ongoing research is the further refine-
ment of molecular mechanics force fields to better de-
scribe the system at hand. It has been shown that both
structural details of simulated bilayers and thermody-
namic properties of membrane translocation vary among
commonly used force fields [66, 67]. However, details of
the force field selection should not alter our primary con-
clusion because the described sampling and convergence
issues are rooted in basic physical properties, specifically
the electrostatic interaction between the solute and the
phospholipid headgroups.
Judging the convergence of free energy calculation
is not an easy task. Because common error analysis
methods may not always reveal characteristics of a non-
converged free energy profile, these calculations need ad-
ditional verification. As useful guidance we propose the
following diagnostic checks:
i. Is the free energy profile symmetric with respect to
the bilayer center? This is a strict requirement for
symmetric membranes.
ii. Does the free energy profile exhibit a plateau re-
gion in the center of the bilayer? Mirrored or sym-
metrized free energy profiles often exhibit a kink at
z = 0. This could be a warning sign for insufficient
convergence, as such a kink would imply a discon-
tinuity in the mean force, which is unlikely to be
physically meaningful.
iii. When performing an UE calculation with initial
conditions taken from simulation trajectories at
varying z-positions of the solute, is there an ap-
parent gap in the resulting exchange pattern? The
presence of such a gap indicates the existence of
9
barriers in replica space that will impede proper
sampling.
iv. For solutes that are most likely found in bulk water
or at the membrane-water interface, do the sam-
pled configurations at the center of the bilayer ex-
hibit a bias towards one side of the membrane over
the other? For symmetric membranes it is likely
that many configurations of the z = 0 ensemble are
transition states, i.e., they have equal probability
of evolving towards states in which the solute is on
either side of the membrane. Calculating the com-
mittor distribution function is an effective if time-
consuming way to test this property: if it is peaked
at pB = 0.5 then one indeed samples the transition
state ensemble, which is not the case if it is peaked
at pB = 0 and/or pB = 1.
None of the three tested methods has an intrinsic ad-
vantage when comparing the amount of sampling that
can be obtained for a fixed amount of computing time.
They differ, however,
in how computational resources
are allocated. WT-MetaD calculations are typically per-
formed as a single, long trajectory. This trajectory can be
run on multiple computing nodes in parallel, but the scal-
ing efficiency is limited by the size of the system. Other
extensions of the metadynamics method, such as Paral-
lel Tempering Metadynamics, allow the parallelization of
multiple metadynamics trajectories in a single calcula-
tion [68]. US, as frequently used and described here, in-
volves performing multiple independent simulations, one
for each biasing window. This can be done either in se-
ries or in parallel depending on the resources available,
which provides the most flexibility. In contrast, general
UE requires that simulations of all replicas are running
in parallel, which necessitates the availability of many
computing cores at the same time. However, there are
variants of the Replica Exchange algorithm, such as Vir-
tual Replica Exchange (VREX) [69] and Serial Replica
Exchange [70] that circumvent the need for synchronous
simulations.
In summary, we believe that the ability to judge con-
vergence by examining the exchange patterns makes UE
an excellent choice if the computing requirements can
be satisfied. It yields robust free energy profiles, which
are essential for gaining a deeper understanding of mem-
brane translocation processes and for predicting perme-
ation rates.
ACKNOWLEDGMENTS
We thank Chris Neale for helpful discussions. This
work was facilitated though the use of advanced compu-
tational, storage, and networking infrastructure provided
by the Hyak supercomputer system at the University of
Washington. We thank the Royalty Research Fund for
partial financial support for this work through Grant 65-
4785.
10
[1] H. D. Herce and A. E. Garcia, Proceedings of the Na-
tional Academy of Sciences of the United States of Amer-
ica 104, 20805 (2007).
[2] Y. Hu and S. Patel, Soft Matter 12, 6716 (2016).
[3] D. P. Tieleman, Clinical and Experimental Pharmacol-
ogy and Physiology 33, 893 (2006).
[4] S. Piantavigna, M. E. Abdelhamid, C. Zhao, X. Qu, G. A.
McCubbin, B. Graham, L. Spiccia, A. P. O'Mullane, and
L. L. Martin, ChemPlusChem 80, 83 (2015).
[5] M. Lindgren, M. Hallbrink, a. Prochiantz, and U. Lan-
gel, Trends in Pharmacological Sciences 21, 99 (2000).
[6] S. Futaki, International Journal of Pharmaceutics 245, 1
(2002).
[30] E. Darve and A. Pohorille, The Journal of Chemical
Physics 115, 9169 (2001).
[31] J. G. Kirkwood, The Journal of Chemical Physics 3, 300
(1935).
[32] A. C. V. Johansson and E. Lindahl, Proceedings of the
National Academy of Sciences of the United States of
America 106, 15684 (2009).
[33] J. L. MacCallum, W. F. D. Bennett, and D. P. Tieleman,
Biophysical Journal 101, 110 (2011).
[34] J. Gumbart, C. Chipot, and K. Schulten, Proceedings of
the National Academy of Sciences of the United States
of America 108, 3596 (2011).
[35] M. A. Wilson and A. Pohorille, The Journal of the Amer-
[7] M. Zorko and U. Langel, Advanced Drug Delivery Re-
ican Chemical Society 118, 6580 (1996).
views 57, 529 (2005).
[8] A. D. Frankel and C. O. Pabo, Cell 55, 1189 (1988).
[9] S. Deshayes, M. C. Morris, G. Divita,
and F. Heitz,
Cellular and Molecular Life Sciences 62, 1839 (2005).
[10] C. Bechara and S. Sagan, FEBS Letters 587, 1693 (2013).
[11] F. Milletti, Drug Discovery Today 17, 850 (2012).
[12] R. J. Clarke, Advances in Colloid and Interface Science
89, 263 (2001).
[13] R. V. Swift and R. E. Amaro, Chemical Biology and Drug
Design 81, 61 (2013).
[14] A. E. Cardenas, R. Shrestha, L. J. Webb, and R. Elber,
The Journal of Physical Chemistry B 119, 6412 (2015).
and
H. Schindler, Molecular Membrane Biology 17, 17 (2000).
[16] C. Ciobanasu, J. Peter Siebrasse, and U. Kubitscheck,
[15] G. J. Schutz, M. Sonnleitner, P. Hinterdorfer,
Biophysical Journal 99, 153 (2010).
[17] A. C. V. Johansson and E. Lindahl, Journal of Chemical
Physics 130, 185101 (2009).
[18] S. Dorairaj and T. W. Allen, Proceedings of the National
Academy of Sciences of the United States of America
104, 4943 (2007).
[19] J. L. MacCallum, W. F. D. Bennett, and D. P. Tieleman,
Biophysical Journal 94, 3393 (2008).
[20] C. L. Wennberg, D. Van Der Spoel, and J. S. Hub, Jour-
nal of the American Chemical Society 134, 5351 (2012).
[21] D. P. Tieleman and S. J. Marrink, Journal of the Amer-
ican Chemical Society 128, 12462 (2006).
[22] C. J. Dickson, V. Hornak, R. A. Pearlstein, and J. S.
Duca, Journal of the American Chemical Society 139,
442 (2017).
[23] C. T. Lee, J. Comer, C. Herndon, N. Leung, A. Pavlova,
R. V. Swift, C. Tung, C. N. Rowley, R. E. Amaro,
C. Chipot, Y. Wang, and J. C. Gumbart, Journal of
Chemical Information and Modeling 314, 721 (2016).
[24] C. Neale, W. F. D. Bennett, D. P. Tieleman,
and
R. Pom´es, Journal of Chemical Theory and Computa-
tion 7, 4175 (2011).
[25] C. Neale, C. Madill, S. Rauscher, and R. Pom´es, Journal
of Chemical Theory and Computation 9, 3686 (2013).
[26] C. Neale and R. Pom`es, BBA - Biomembranes 1858,
2539 (2016).
[27] A. Barducci, G. Bussi, and M. Parrinello, Physical Re-
view Letters 100, 20603 (2008).
[28] G. M. Torrie and J. P. Valleau, Journal of Computational
Physics 23, 187 (1977).
[29] Y. Sugita and Y. Okamoto, Chemical Physics Letters
314, 141 (1999).
[36] N. Sapay, W. F. D. Bennett, and D. P. Tieleman, Soft
Matter 5, 3295 (2009).
[37] M. Orsi and J. W. Essex, in Molecular Simulations and
Biomembranes : From Biophysics to Function (Royal So-
ciety of Chemistry 2010, 2010) 1st ed., pp. 76–90.
[38] J. Comer, K. Schulten, and C. Chipot, Journal of Chem-
ical Theory and Computation 10, 554 (2014).
[39] W. C. Wimley and S. H. White, Nature Structural Biol-
ogy 3, 842 (1996).
[40] E. Awoonor-Williams and C. N. Rowley, BBA - Biomem-
branes 1858, 1672 (2016).
[41] D. Van Der Spoel, E. Lindahl, B. Hess, G. Groenhof,
A. E. Mark, and H. J. C. Berendsen, Journal of Com-
putational Chemistry 26, 1701 (2005).
[42] M. Bonomi, D. Branduardi, G. Bussi, C. Camil-
loni, D. Provasi, P. Raiteri, D. Donadio, F. Marinelli,
F. Pietrucci, R. A. Broglia, and M. Parrinello, Com-
puter Physics Communications 180, 1961 (2009).
[43] T. Darden, D. York, and L. Pedersen, The Journal of
Chemical Physics 98, 10089 (1993).
[44] B. Hess, H. Bekker, H. J. C. Berendsen, and J. G. E. M.
Fraaije, Journal of Computational Chemistry 18, 1463
(1997).
[45] O. Berger, O. Edholm, and F. Jahnig, Biophysical Jour-
nal 72, 2002 (1997).
[46] H. J. C. Berendsen, J. P. M. Postma, W. F. V. Gunsteren,
and J. Hermans, Intermolecular Forces , 331 (1981).
[47] G. A. Kaminski, R. A. Friesner, J. Tirado-Rives, and
W. L. Jorgensen, The Journal of Physical Chemistry B
105, 6474 (2001).
[48] A. C. V. Johansson and E. Lindahl, Proteins 70, 1332
(2008).
[49] W. Humphrey, A. Dalke, and K. Schulten, Journal of
Molecular Graphics 14, 27 (1996).
[50] A. Laio and F. L. Gervasio, Reports on Progress in
Physics 71, 126601 (2008).
[51] J. F. Dama, M. Parrinello, and G. A. Voth, Physical
Review Letters 112, 240602 (2014).
[52] D. Bochicchio, E. Panizon, R. Ferrando, L. Monticelli,
and G. Rossi, The Journal of Chemical Physics 143,
144108 (2015).
[53] S. Kumar, J. M. Rosenberg, D. Bouzida, R. H. Swendsen,
and P. A. Kollman, Journal of Computational Chemistry
13, 1011 (1992).
[54] J. S. Hub, B. L. D. Groot, and D. V. D. Spoel, Journal
of Chemical Theory and Computation 6, 3713 (2010).
11
[55] K. Hukushima and K. Nemoto, Journal of the Physical
Society of Japan 65, 1604 (1996).
[56] P. G. Bolhuis, D. Chandler, C. Dellago,
and P. L.
Geissler, Annual Review of Physical Chemistry 53, 291
(2002).
[57] C. Dellago, P. G. Bolhuis, and P. L. Geissler, Advances
in Chemical Physics, January 2002 (2001) pp. 1–78.
[58] T. Hessa, H. Kim, K. Bihlmaier, C. Lundin, J. Boekel,
H. Andersson, I. Nilsson, S. H. White, and G. von Heijne,
Nature 433, 377 (2005).
[59] I. Vorobyov, T. E. Olson, J. H. Kim, R. E. Koeppe, O. S.
Andersen, and T. W. Allen, Biophysical Journal 106,
586 (2014).
[60] S. Yesylevskyy, S. J. Marrink, and A. E. Mark, Biophys-
ical Journal 97, 40 (2009).
[61] C. Neale, J. C. Y. Hsu, C. M. Yip, and R. Pom´es, Bio-
physical Journal 106, 29 (2014).
[62] W. F. D. Bennett, C. K. Hong, Y. Wang, and D. P.
Tieleman, Journal of Chemical Theory and Computation
12, 4524 (2016).
[63] D. Lin and A. Grossfield, Biophysical Journal 109, 750
(2015).
[64] M. J. Hinner, S. J. Marrink, and A. H. D. Vries, The
Journal of Physical Chemistry B 113.
[65] Y. Wang, D. Hu, and D. Wei, Journal of Chemical The-
ory and Computation 10, 1717 (2014).
[66] T. J. Piggot, , A. Pinero,
and S. Khalid, Journal of
Chemical Theory and Computation 8, 4593 (2012).
[67] D. Sun, J. Forsman, and C. E. Woodward, Journal of
Chemical Theory and Computation 11, 1775 (2015).
[68] G. Bussi, F. L. Gervasio, A. Laio, and M. Parrinello,
Journal of the American Chemical Society 128, 13435
(2006).
[69] S. Rauscher, C. Neale, and R. Pom´es, Journal of Chem-
ical Theory and Computation 5, 2640 (2009).
[70] M. Hagen, B. Kim, P. Liu, R. A. Friesner, and B. J.
Berne, Journal of Physical Chemistry B 111, 1416
(2007).
Supplemental Information
Free Energy Calculations of Membrane Permeation:
Challenges due to Strong Headgroup–Solute Interactions
1
TABLE S1. Metadynamics Parameters
Solute
Height(kJ/mol) Width(nm) Time of gaussian addition(ns) Biasfactor
15
Zwitterionic arginine
Sodium
5.0
Side chain analog of alanine 0.1
Alanine with neutral termini 10
15
Zwitterionic alanine
Water
2.2
140
75.0
2.00
35.0
140
10.0
0.3
0.3
0.1
0.3
0.3
0.1
3
3
5
3
3
3
TABLE S2. Lengths of Simulation Trajectories
US
UE
length/window (ns) total (µs) length/replica (ns) total ∗ (µs)
100
-
WT-metad (µs)
∗
UE total simulation time only represent simualtions that span the membrane between −2.0nm< z ≤ 2.nm. We ran
additional simulation around z− ≤ 2.0 nm and z > 2.0 nm to investigate if solute configuration changes in bulk water
phase.
Solute
Zwitterionic arginine
Sodium
Side chain analog of alanine 20
Alanine with neutral termini 20
Zwitterionic alanine
55
-
Water
4.00
-
1.00
1.00
2.75
-
166
27
-
-
55
8
6.64
1.08
-
-
2.20
0.32
1.00
1.08
1.00
1.00
1.40
0.42
2
FIG. S1. UE results for sodium. (a) Exchange pattern lines for 40 replicas. Replicas starting at z = 2.0 nm, 0 nm and −1.9 nm are
highlighted. A gap in the exchange pattern (indicated by the arrow) is clearly visible. Using the final configuration of z = 0 replica as the
initial condition for the center replicas in a new UE simulation (b), this gap is no longer present, indicating even exchanges between all 40
replicas.
Replicacenter[nm]SimulationTime[ns](a)gapReplicacenter[nm]SimulationTime[ns](b)-101201020304050607080-101205101520253
FIG. S2. UE results for zwitterionic form of alanine. (a) Exchange pattern lines for 40 replicas. Replicas starting at z = 2.0 nm, 0 nm
and −1.9 nm highlighted. A gap in the exchange pattern (indicated by the arrow) is clearly visible. Using the final configuration of z = 0
replica as the initial condition for the center replicas in a new UE simulation (b), this gap is no longer present, indicating even exchanges
between all 40 replicas.
Replicacenter[nm]SimulationTime[ns](a)gapReplicacenter[nm]SimulationTime[ns](b)-1012010203040506070-1012010203040504
FIG. S3. UE results for water. Exchange pattern lines for 40 replicas. Replicas starting at z = 2.0 nm, 0 nm and −1.9 nm highlighted.
No gap in the exchange pattern exists indicating an even exchanges between all 40 replicas.
Replicacenter[nm]SimulationTime[ns]-101202468 |
1411.2782 | 3 | 1411 | 2015-10-18T18:18:12 | Forced desorption of semiflexible polymers, adsorbed and driven by molecular motors | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech",
"q-bio.SC"
] | We formulate and characterize a model to describe the dynamics of semiflexible polymers in the presence of activity due to motor proteins attached irreversibly to a substrate, and a transverse pulling force acting on one end of the filament. The stochastic binding-unbinding of the motor proteins and their ability to move along the polymer, generates active forces. As the pulling force reaches a threshold value, the polymer eventually desorbs from the substrate. Performing molecular dynamics simulations of the polymer in presence of a Langevin heat bath, and stochastic motor activity, we obtain desorption phase diagrams. The correlation time for fluctuations in desorbed fraction increases as one approaches complete desorption, captured quantitatively by a power law spectral density. We present theoretical analysis of the phase diagram using mean field approximations in the weakly bending limit of the polymer and performing linear stability analysis. This predicts increase in the desorption force with the polymer bending rigidity, active velocity and processivity of the motor proteins to capture the main features of the simulation results. | physics.bio-ph | physics | Forced desorption of semiflexible polymers, adsorbed and driven by molecular motors
Abhishek Chaudhuri1, ∗ and Debasish Chaudhuri2, †
1 Indian Institute of Science Education and Research Mohali,
Knowledge City, Sector 81, SAS Nagar - 140306, Punjab, India.
2Indian Institute of Technology Hyderabad, Yeddumailaram 502205, Telangana, India
(Dated: November 6, 2018)
We formulate and characterize a model to describe dynamics of semiflexible polymers in the pres-
ence of activity due to motor proteins attached irreversibly to a substrate, and a transverse pulling
force acting on one end of the filament. The stochastic binding-unbinding of the motor proteins
and their ability to move along the polymer, generates active forces. As the pulling force reaches a
threshold value, the polymer eventually desorbs from the substrate. Performing molecular dynam-
ics simulations of the polymer in presence of a Langevin heat bath, and stochastic motor activity,
we obtain desorption phase diagrams. The correlation time for fluctuations in desorbed fraction
increases as one approaches complete desorption, captured quantitatively by a power law spectral
density. We present theoretical analysis of the phase diagram using mean field approximations in
the weakly bending limit of the polymer and performing linear stability analysis. This predicts
increase in the desorption force with the polymer bending rigidity, active velocity and processivity
of the motor proteins to capture the main features of the simulation results.
PACS numbers: 05.20.-y, 36.20.-r, 87.15.-v
I.
INTRODUCTION
Cytoskeleton in the cell comprises of semiflexible pro-
tein filaments, cross-linkers and motor-proteins, and is
maintained continuously out of equilibrium. Each fam-
ily of motor proteins, when coupled to their type-specific
filamentous tracks, can hydrolyze chemical fuel (ATP),
generating motion and stresses in the cell [1 -- 4]. This
active meshwork provides the cell its mechanical stabil-
ity [5, 6], tracks for intra-cellular locomotion, control-
ling cell-motility [7, 8], as well as organizing signalling
platforms on the cell membrane and endocytosis [9, 10].
Single molecule experiments on motor proteins revealed
mechanism of force generation, force-velocity relations,
and dependence of motion on ATP concentration [11, 12].
Collective action of molecular motors lead to interesting
dynamics like bidirectional motion and spontaneous os-
cillations [13 -- 16]. The transport of cargo in one dimen-
sion (1D) by multiple motors has also attracted much
attention and the response to external opposing forces
have been obtained [17 -- 19].
A plethora of individual and collective physical prop-
erties of cytoskeletal filaments were obtained from the
study of in vitro gliding assays,
in which F-actins or
microtubules move on a two dimensional substrate dec-
orated by myosin or kinesin motors, both experimen-
tally [20 -- 25] and theoretically [26 -- 29, 29, 30]. Recent ex-
periments on molecular motor assays revealed formation
of spiral defects and loops of actively moving filaments
driven by motor proteins [22, 31]. In gliding assays, one
end of motor proteins are irreversibly attached to a two
∗Electronic address: [email protected]
†Electronic address: [email protected]
dimensional substrate. The other end binds to the fila-
ments and actively forces them to move parallel to the
substrate. In this context, it is important to understand
the response to external forces of these gliding filaments
actively driven by the molecular motors.
In this paper, we consider a semiflexible filament float-
ing on a molecular motor assay in presence of a pulling
force acting on one end of the filament in a direction
perpendicular to the substrate. The filament is actively
captured and driven parallel to the substrate by molecu-
lar motors. The situation is akin to in vivo microtubules,
one end of which is captured and actively driven by motor
proteins at the cell cortex while forces act on the other
end attached to the microtubule originating centre [32].
A passive counter-part of this problem is peeling of semi-
flexible polymers from adhesive surfaces [33 -- 39]. Such
peeling experiments have been shown to be important
in quantifying the strengths of actomyosin rigor bonds,
in absence of ATP driven activity, measured by pulling
F-actins off myosin coated substrate using optical tweez-
ers [40]. Rupture of multiple bonds have also been stud-
ied in the context of cell adhesion [41 -- 45] and the unzip-
ping of DNA [46 -- 48]. Semiflexible polymers themselves
are known to show interesting mechanical and dynamic
properties [49 -- 55].
We use Langevin dynamics simulation and theortical
mean field analysis to quantify the dynamics, and hence
the response, of the semiflexible filament and the molec-
ular motors, under the transverse pulling force. For sim-
plicity, we assume that the motor proteins, which un-
dergo attachment-detachment kinetics, are arranged uni-
formly on a one-dimensional substrate. With increase
in the transverse pulling force applied to one end of the
polymer, it undergoes a non-equilibrium continuous tran-
sition at a threshold force, from an adsorbed state to a
completely desorbed state. A theorteical mean field anal-
5
1
0
2
t
c
O
8
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
3
v
2
8
7
2
.
1
1
4
1
:
v
i
X
r
a
ysis of the problem predicts an increase in the threshold
force of desorption with increasing bending rigidity of
polymer, as well as activity and processivity of molecu-
lar motors. The predictions show good agreement with
simulation results. We obtain results for biologically rele-
vant parameter values, making our predictions amenable
to direct verification in gliding assay experiments.
II. MODEL
We model the cytoskeletal filaments as stretchable
semiflexible polymers described by space curve r(s), and
local tangent vector t(s) = ∂r/∂s with s denoting a po-
sition along the contour of the chain. The Hamiltonian
for a filament of length L is given by [56]
(cid:19)2(cid:35)
(cid:18) ∂r
∂s
Parameters
Definition
Values
γw
T
α
σ
A
km
ρ
κAT
κMT
v0
ωon
ω0
fd
fs
Viscosity of water
Temperature
Viscous drag
Bond length
0.001 pN-s/µm2
4.2 pN-nm/kB
1 pN-s/µm
0.5 µm
100 pN/µm
100 pN/µm
Spring constant(filament)
Spring constant(motor)
Linear density (motor)
0.07 pN-µm2
Bending rigidity (Actin)
Bending rigidity (µtubule) 21.84 pN-µm2
0.8 µm/s (K)
Free motor velocity
2.5/µm
Attachment rate
Bare detachment rate
Detachment force
20/s (K)
1/s (K)
6 pN (K)
Stall force
1.16 pN (K)
(cid:90) L
(cid:34)
(cid:18) ∂2r
(cid:19)2
H =
1
2
ds
κ
0
∂s2
+ A
TABLE I: Various parameters and their typical values used
in the simulation. (K) denotes kinesin.
(1)
where κ is the bending rigidity, A is the bond strength.
This model reduces to the worm like chain model in the
limit of unstretchable bonds with [t(s)]2 = 1.
0 = (xi
The motor proteins are modeled as elastic linkers. The
tail end of i-th motor protein is attached irreversibly to
the substrate at position ri
0, 0). The head end is
free to attach (detach) to (from) the filament. It attaches
to a segment of the filament if it lies within a capture ra-
dius rc with an attachment rate ωon. When attached,
the motor head either moves along the filament or gets
detached from it with a rate ωoff . The i-th molecular
motor when attached to the polymer at a position r(s)
exerts an elastic force fm = −km(r(s) − ri
0) = −kmδr,
attracting the polymer segment towards itself. The am-
plitude of this force is the load fl = fm on the head of a
molecular motor. In their active state, the attached end
of motor proteins move along the filament towards one of
its ends (plus end for kinesins walking on microtubules)
with active relative velocity [see Fig.1(a)]
va
t (ft) =
−v0
1 + d0 exp(ft/fs)
(2)
where ft = −fm.t, d0 = 0.01 and fs = 1.16 pN is the stall
force (parameters corresponding to kinesin molecule, see
Appendix A). The negative sign is chosen so that the
motors move towards the s = 0 end of the polymer. This
ATP-driven motion of the motor heads, via their elastic
nature generate an active force on the filament in a di-
rection opposite to this movement, resulting in a sliding
motion of the filament with respect to the substrate. The
external force Fz, applied at the s = 0 end of the poly-
mer opposes this active force, and at sufficient strength
desorbs the polymer from the substrate.
Note that, ωon and ωoff are dependent on the separa-
tion of a given polymer segment from the substrate. In
our simplified quasi one-dimensional model, the motors
are attached uniformly with the density ρ along the x-
axis and the transverse fluctuations of the polymer and
the motor heads are along the z−axis (Fig. 1(a)). As-
suming a Kramer's process we have, ωoff = ω0 exp(fl/fd),
where ω0 is the bare off rate, fl is the load force origi-
nating from the elastic extension of the motor spring and
fd is the typical force required to detach the motor head
from the polymer segment. An external force Fz, applied
at the s = 0 end of the polymer opposes this active force,
and at sufficient strength desorbs the polymer from the
substrate.
III. SIMULATION
(cid:80)N−1
n=1 (κ/2σ)[t(n + 1) − t(n)]2 +(cid:80)N
To study the full dynamics of the semiflexible poly-
mer under the influence of motor proteins attached to a
substrate, and pulled out of the substrate by an exter-
nal force, we perform molecular dynamics (MD) simu-
lations of the polymer in presence of a Langevin heat
bath, and stochastic attachment detachment kinemat-
ics of molecular motors.
In simulations, we discretize
the semiflexible polymer into a bead-spring chain of N
bonds of equilibrium length σ, spring constant A, and fi-
nite bending rigidity κ such that the Hamiltonian is H =
n=1(A/2)[b(n) − σ]2.
Here we denoted position of the n-th bead by r(n), such
that the local tangent t(n) = [r(n + 1) − r(n)]/b(n)
where b(n) = r(n + 1) − r(n)
is the instantaneous
bond length.
instantaneous
bond lengths b(n) ≈ σ, and the chain in equilibrium
In addition, we incor-
behaves like a worm like chain.
porate self avoidance via a Weeks-Chandler-Anderson
(WCA) purely repulsive potential between non-bonded
polymer beads βVW CA(rij) = 4[(σ/rij)12−(σ/rij)6+1/4]
if rij < 21/6σ and 0 otherwise, with β = 1/kBT the in-
verse temperature.
In the limit of large A,
The MD simulations are performed using a velocity-
Verlet algorithm in presence of a Langevin heat bath
2
FIG. 1: (Color online) Polymer configuration, and adsorption-desorption transitions. Units used are: positions in µm, forces
in pN, velocities in µm/s. (a) A configuration of semiflexible bead-spring polymer (red beads and lines) actively captured
by motor proteins on a substrate. Each motor protein is denoted by a blue line connecting two end-points, one of which is
bound irreversibly to the substrate while the other end can attach (detach) to (from) the polymer beads with fixed rates.
The end of motor proteins attached to a polymer bead walks towards the trailing end of the polymer on which an external
force Fz is applied. This in turn pulls the filament in the opposite direction. (Inset) The time averaged desorbed fraction of
polymer length (cid:104)ε(cid:105) as a function of Fz, where ε = (cid:96)/L with (cid:96) denoting desorbed length of the polymer. (cid:104)ε(cid:105) remains small up to
17 pN, subsequently showing gradual increase finally desorbing completely at a threshold force F c
z = 22pN. (b) Phase diagrams
depicting adsorption-desorption transition of a stiff (microtubule) and a relatively flexible (F-actin) filament, as a function of
active velocity v0 of motor proteins. Both of them show increase in desorption force with increasing v0, finally merging into a
single curve at high enough values of v0. Lines are fit to Eq. 5 with b(cid:48)
1 = 1/µm, b(cid:48)
3 = 18.53
pN/µm2 for F-actin and b(cid:48)
3 = 0.755 pN/µm2 for microtubule. (c) Phase
diagram for Fz vs Ωd for v0 = 0.807µm/s for the same two filaments as in (b). The line is a fit to Eq. 6 with the same b(cid:48)
0 as
obtained in (b) for microtubule.
0 = 0.56, b(cid:48)
2 = 15.126 pN-s/µm3 and b(cid:48)
2 = 5.11 pN-s/µm3 and b(cid:48)
0 = 0.384, b(cid:48)
1 = 1/µm, b(cid:48)
that fixes the temperature at the room temperature
value kBT = 4.2 pN-nm through a Gaussian white noise
(cid:104)ηi(t)ηj(t(cid:48))(cid:105) = 2αkBT δijδ(t − t(cid:48)) with α denoting viscos-
ity of the environment. Since the typical environment
within a cell is at least one order of magnitude more vis-
cous than water, we choose the viscosity of the medium
γ = 100γw = 0.1 pN s/µm2, where γw = 0.001 pN s/µm2
is the viscosity of water. Therefore, the viscous drag
α = 6πγa ≈ 1 pN s/µm on a bond of length σ = 0.5 µm.
The spring constant of the bead-spring system is taken to
be fairly large A = 100 pN/µm so that the bond fluctu-
ations are small enough to reproduce known equilibrium
statistics [49, 53]. The persistence length λ of cytoskele-
tal filaments varies by three orders of magnitude, with
λ = 16.7 µm for actin filaments, to λ = 5.2 mm for mi-
crotubules [57]. These correspond to variation of bending
rigidity κ from the value κAT = 0.07 pN-µm2 for F-actins
to κMT = 21.84 pN-µm2 for microtubules. Unless stated
otherwise, in our simulations we consider parameters typ-
ical for kinesin motors with attachment rate, ωon = 20/s
and a bare detachment rate ω0 = 1/s. The motors are
placed on a 1d line along the x-axis with constant cover-
age density. The spring constant, km, for kinesin motors
lie between 10-1000 pN/µm. In our simulations, we use
km = A = 100 pN µm−1. The detachment force, fd = 6
pN characterizes the force induced enhancement of de-
tachment rates as ωoff = ω0 exp(fl/fd) where fl is the
instantaneous load on the molecular motor. In our sim-
ulations, we used the polymer bond-length σ as unit of
length, and the typical forces associated with the motor
proteins 1 pN as the unit of force.
The unit of time is set by tu = α/A, and we choose
the integration time step δt = 0.01tu. Attachment- de-
tachment kinematics of motor proteins are performed
stochastically with probabilities ωonδt and ωoff δt at ev-
ery time-step. Note that the attachment event is tried
only if a filament segment is within the capture radius
rc ∼ σ, from the equilibrium position of molecular mo-
tors. Once detached the molecular motors are assumed
to relax back to equilibrium configurations immediately.
When attached to the polymer, molecular motors move
along the polymer following Eq.(2).
In the absence of
load, the attached motors move with velocity v0 in the
negative x direction along the filament, forcing the poly-
mer to translate towards the positive x direction.
To study the effect of transverse pulling force, Fz is
applied in the z direction perpendicular to the substrate
and at the trailing end of the polymer (see Fig. 1 (a)).
We study the influence of the pulling force Fz as we vary
(i) active self-propulsion v0, (ii) the duty ratio Ωd, and
(iii) the bending rigidity κ. All the parameters used in
the simulations are summarized in Table I.
IV. RESULTS
In the absence of a transverse pulling force and for
processive motors with large duty ratio Ωd = 0.95 (ωon :
ω0 = 20 : 1), desorption of the polymer by stochastic
fluctuations is prevented. The polymer stays on the mo-
3
00.20.40.60.81−10−5051015(a)zxFz00.5110141822hεiFzFz01020304000.20.40.60.81(b)Fzv0AdbedDebed051015200.50.60.70.80.91(c)FzΩdAdbedDebedκATκTκATκTFIG. 2: (Color online) Phase diagram for Fz (in pN) as a
function of κ (in pN-µm2) for a passive system (v0 = 0, (cid:4)),
intermediate activity (v0 = 0.2µm/s, (cid:32)) and high activity
(v0 = 0.807µm/s, (cid:78)). Lines are fit to Eq. 5 with b(cid:48)
b(cid:48)
1 = 1/µm, b(cid:48)
3 = 20.97 pN/µm2 for v0 = 0, b(cid:48)
1 = 1/µm, b(cid:48)
2 = 28.737 pN-s/µm3, b(cid:48)
b(cid:48)
v0 = 0.2 and b(cid:48)
0 = 0.33, b(cid:48)
1 = 1/µm, b(cid:48)
b(cid:48)
3 = 20.26 pN/µm2 for v0 = 0.807.
0 = 0.267,
0 = 0.278,
3 = 20.97 pN/µm2 for
2 = 28.737 pN-s/µm3,
tor protein bed and slides towards positive x-axis with
a velocity close to v0 characteristic of individual motor
proteins. The response of the system to a force applied
parallel to the substrate is characterized in detail in Ap-
pendix B. In this specific case, filament bending does not
play any role and the filament is well approximated as a
one dimensional rigid rod.
For a transverse pulling force applied on the trailing
end, opposing the sliding motion of the filament, we
study the transition of the polymer from an adsorbed
to a completely desorbed state. Fig.1(a) shows a typical
configuration of a filament having κ = κMT the bend-
ing rigidity of microtubules, on a bed of motor proteins
having active velocity v0 = 0.807 µm s−1 corresponding
to kinesins, under external force Fz. For further details
on modeling kinesin activity see Appendix A. The des-
orption is characterized by a continuous increase in the
fraction of desorbed length beyond a threshold force, as
shown in the inset of Fig. 1(a), or by following the frac-
tion of motor proteins attached to the filament. Note
that beyond Fz = 17pN in Fig. 1(a), the polymer starts
to partially desorb, while the complete desorption oc-
curs at a relatively higher pulling force of F c
z = 22pN.
The transition from adsorbed to desorbed state occurs
smoothly, like a continuous phase transition. We come
back to this point again at the end of this section. We
obtain F c
z with changing v0, Ωd and κ, in each case keep-
ing the other parameter values unchanged. This gives
us three phase diagrams for adsorption-desorption tran-
sition.
Fig. 1(b) shows the dependence of F c
z on v0, for two dif-
ferent values of bending rigidity of polymer κ: κ = κAT
corresponds to relatively flexible F-actins, and κ = κMT
FIG. 3: (Color online) Time series of the desorbed fraction
of microtubule from kinesin bed, at different values of pulling
force Fz expressed in units of pN. A longer time series is
presented at Fz = 20.1pN, a force value close to the desorption
transition.
corresponds to the very rigid limit of microtubules. The
simulation results show that both for F-actins and mi-
crotubules the threshold desorption force F c
z increases
with v0, to eventually merge together and saturate at
large values of v0 in accordance with Eq. 5, derived in
the following section. Both the data sets in Fig. 1(b)
fit well with Eq. 5. Fitting parameters are mentioned
in the figure captions. F c
z increases with v0 to eventu-
z ∼ [1 − C1/(C2 + v0)]1/2,
ally saturate at large v0 as F c
where C1, C2 are constants. At large enough values of v0
and κ, the desorption force is expected to become inde-
pendent of both [see Eq.6], leading to the same F c
z for
F-actins and microtubules for large active velocity v0 of
the molecular motors. Thus the two phase boundaries
for F-actins and microtubules merge together as v0 ap-
proaches 1µm/s. Fig. 1(c) presents the simulated phase
diagram in F c
z − Ωd plane, calculated with large active
motion v0 = 0.807 µm/s. It shows good agreement with
mean field results (Eq. 6).
In Fig. 2, we present the dependence of desorption force
F c
z on bending stiffness κ of the polymer, for different val-
ues of active velocity v0. All the data sets fit well to Eq. 5.
A passive semi-flexible polymer (v0 = 0) shows increase
(cid:48)
2 + κ)]1/2,
in F c
(cid:48)
with C
2 are constants. This scenario is equivalent
to equilibrium desorption of semiflexible filaments from
adhesive substrates [34, 35]. However, at large values
of active velocity of motor-proteins (v0 = 0.807 µm/s),
v0 would dominate over κ, giving rise to an essentially
κ-independent desorption force.
z with increasing κ as F c
(cid:48)
1 and C
z ∼ [1 − C
(cid:48)
1/(C
Finally we take a closer look at the desorption process
itself. For this purpose we use the time evolution of the
order parameter, the desorbed fraction of the polymer,
ε. The time series of this stochastic quantity ε(t) shows
different behavior depending on the value of the applied
4
051015202530048121620Fzκ0.110.1110102103104εt(s)Fz=17Fz=18Fz=20.1up a tension along the chain. In the following, we assume
an unstretchable chain in which the tension alone bears
the impact of these two opposing forces.
A. Mean field analysis
In the mean field limit, let us assume that nb,u denote
the local density of bound and unbound motors, such
that nb + nu = ρ is a constant. In the detached state mo-
tors play no role in the dynamics of filament. However,
the detached motor heads after performing diffusive re-
laxation reattaches to a polymer segment within capture
radius with a given rate. We assume that the attachment
of a motor head to the polymer is not instantaneous but
happens over a time ta. Further, to attach to the fil-
ament, the motor head diffuses over a distance z, the
local transverse distance of the polymer segment, with
a time scale z2/Du. Therefore, the total time required
for the process is (ta + z2/Du), giving rise to an effective
attachment rate, ωon = Du/(taDu + z2). Thus the dy-
namics of unbound motors may be incorporated in terms
of the effective ωon. The bound motors generate elastic
force per unit length −kmz nb on the filament, and slides
the filament with a velocity −va
t (ft) where ft ∝ Fz/nbL.
The bound motors detach from the filament with
rate ωoff = ω0 exp(fl/fd), where the local load force,
fl = −kmδr. This stochastic load originates from three
mechanisms:
(i) stochastic binding/unbinding of other
motor proteins changing the number of bound motors
sharing the load, (ii) motion of bound head of motor
proteins along the filament, and finally (iii) the external
force, Fz, acting at the filament end.
In the absence
of Fz, stochastic binding/unbinding will result in an av-
erage time-independent separation δr and the sliding
motion of the filament will hand over binding from one
motor to its neighbor without impacting the polymer dy-
namics on an average. Therefore, the average load force
would really be due to the external force Fz. Within
mean-field approximation, we assume that Fz acting on
the polymer is distributed equally among all bound mo-
tors. Therefore, one may use fl = Fz/nbL to obtain
ωoff = ω0 exp(f /nbfd), where f = Fz/L. This remains a
good approximation within the weakly bending limit.
In the limit of small transverse displacements z(x, t) of
the filament from x-axis, the Hamiltonian (Eq. 1) may
be approximated as
(cid:90) L
(cid:34)
(cid:18) ∂2z
(cid:19)2
(cid:19)2(cid:35)
(cid:18) ∂z
∂x
H =
1
2
ds
κ
0
∂x2
+ τ (x, t)
,
(3)
where instead of a large spring constant A used in the
simulations, we use a local instantaneous tension τ (x, t)
in the theory, with τ (x, t) constraining the local bond
lengths to a constant value σ. The over-damped mo-
tion due to this Hamiltonian is described by α⊥∂tz =
−δH/δz + η(t), where η(t) is a Gaussian white noise,
and α⊥ viscous friction. Averaging over the stochas-
5
FIG. 4: (Color online) Power spectrum of the desorbed frac-
tion ε(ν)2 obtained from the respective time series as shown
in Fig. 3. The data show power law profiles 1/να with α ≈ 1
at Fz = 17pN, and α ≈ 2 at Fz = 20.1pN, and reflect an
increase in the correlation time as one approaches the desorp-
tion transition force Fz = 22pN.
desorbing force Fz (Fig.3). At long time ε(t) reaches a
steady state, with a time-independent mean value. The
gradual approach to final steady states with ε ≈ 0.4 for
Fz = 18pN to ε ≈ 0.6 for Fz = 20.1pN shows how the
polymer desorption progresses in time. Note that the
continuous change in average order parameter (cid:104)ε(cid:105) with Fz
as shown in Fig. 1(a) inset, indicates a continuous non-
equilibrium transition. The power spectrums presented
in Fig. 4 quantifies the dynamics. This clearly shows
that the correlation time of the dynamics increases as one
increases Fz towards desorption transition value. The
power spectral density shows a 1/να behavior with α ≈ 1
for Fz ≤ 17pN, and α ≈ 2 as one approaches desorption,
e.g., see the behavior at Fz = 20.1pN. This is reminiscent
of pink noise observed in diverse types of non-equilibrium
systems [58, 59].
V. THEORY
The full dynamics is fairly complicated and not
tractable analytically. However, simplifications of the
original problem is useful to develop analytical insight.
In the following section, we use small bending approx-
imation for the filament, and mean field approximation
for dynamics and external force sharing. Further, the net
impact of the molecular motors on the filament is treated
in terms of a local elastic binding to the substrate, and
a sliding velocity. Since the transverse force is applied
at the end of the polymer opposite to the direction of
sliding, a component of the external force acts against
the direction of sliding. In the simulations, these oppos-
ing forces result partially into elastic energy stored in the
extension of stretchable but stiff bonds, and into building
10−910−810−710−610−510−410−310−210−410−310−210−11ε(ν)2ν(Hz)Fz=17Fz=18Fz=20.1tic noise, and incorporating the force due to bound mo-
tors, the evolution of the transverse displacement and the
attachment-detachment dynamics of the unbound motors
is given by,
α⊥ ∂z
∂t
∂nu
∂t
∂4z
∂x4 + τ (x, t)
= −κ
= ωoff ρ − (ωon + ωoff )nu.
∂2z
∂x2 − kmz nb
(4)
The tension τ (x, t) needs to be determined using the in-
extensibility constraint.
In the weakly bending limit,
spatial variation in τ can be neglected [52], considerably
simplifying the analysis. For values of the external force,
Fz, less than the critical desorption force, the polymer
reaches a steady state configuration z(x) (independent
of t) where it is partially adsorbed.
B. Linear Response
force Fz.
To get an estimate of the critical
force required
to desorb the filament from the substrate, we per-
form a linear stability analysis by assuming a steady
state configuration [z(x), nb(x)] obtained under a fixed
external
Let us consider small variations
[δz(x), δnb(x)] about it. As stated earlier, the attachment
rate ωon = D/(taD + z2), and detachment rate ωoff =
ω0 exp(f /nbfd). Small variations around steady state
give δωoff = −(ωoff f /n2
bfd)δnb, δωon = −[2ωonz/(taD +
z2)]δz. In practice, a motor protein may attach to a seg-
ment of filament only if it lies within a capture radius rc.
We assume that the segment of filament to which mo-
tors may attach remains essentially parallel to the sub-
strate and within a separation σ. Thus replacing rc by
σ, kmz δnu by kmσ δnu, and rewriting δωon = −b1ωonδz
with b1 = 2σ/(taD + σ2), and δnb by −δnu we obtain
from Eq.(4),
(cid:20) ωoff f
α⊥∂tδz = (−κ∂4
x + τ ∂2
∂tδnu =
nbfd − (ωon + ωoff )
x − kmnb)δz + kmσ δnu
δnu + b1ωonnuδz.
(cid:21)
As argued before, the total tensile force τ may be ex-
pressed in terms of active processive motion of the motor
proteins as τ = b2va
t + b3, with b2 a undetermined con-
stant. Here the constant b3 denotes the tension due to
joint action of external force Fz and adhesion to substrate
by the motor proteins. The linear perturbations consid-
ered above are variations around a steady state where we
assumed that all tension propagation [52] is settled down.
Thus a linear dependence of τ on a steady state velocity
va
t (Fz) is reasonable.
If one considers a small segment of a long polymer,
far away from the boundary on which pulling force Fz is
applied, boundary conditions would not affect the local
behavior. In this limit, performing a Fourier transform,
the evolution of specific modes follows ∂t(δzq, δnq
u) =
A.(δzq, δnq
u), where the elements of the 2 × 2 matrix A
6
(cid:104) ωoff f
(cid:105)
.
nbfd − (ωon + ωoff )
are given by A11 = −[κq4 + τ q2 + kmnb], A12 = kmσ,
A21 = b1ωonnu, and A22 =
In the large q limit,
the unstable modes of the
above linearized dynamics identifies the condition where
the absorbed state of the polymer is locally unsta-
ble. Thus it identifies an upper bound of
instabil-
ity, which is instructive to study. However, the ac-
tual desorption may take place at a smaller force.
In
(cid:112)
the large q limit, we have A11 ≈ −[κq4 + τ q2]. The
two eigenvalues of matrix A are λ± = 1
2 [A11 + A22 ±
(A11 + A22)2 − 4(A11A22 − A12A21)]. The condition
that λ+ > 0 for the mode to be unstable is satisfied
if A12A21 − A11A22 > 0.
To obtain a closed analytic form for the expression
of instability condition, we linearize the force depen-
dence of detachment rate ωoff ≈ ω0(1 + Fz/Nbfd), where
Nb = nbL. This assumption is reasonable in the weakly
bending limit where a large number of motor proteins
would be in the bound state. Note that in the absence of
external force, bound motor density n0
b = ρΩd, with the
duty ratio of motor proteins, Ωd = ωon/[ωon + ω0]. As-
suming σ as the smallest length scale in the problem, so
that q ∼ 1/σ, the condition λ+ > 0 leads to an inequal-
ity identifying a force which will destabilize the steady
state profile of an adsorbed filament. Thus the following
expression gives the critical desorption force,
(cid:21)1/2
Ωdkm nuσ4
2v0 + b(cid:48)
κ + b(cid:48)
3
,
(5)
(cid:20)
F c
z ≈
(cid:48)
1 − b
1
b(cid:48)
0fdNb
√1 − Ωd
2 = σ2b2 and b(cid:48)
using vt ≈ v0 in the limit of small Fz/Nb (cid:28) fs. Here
b(cid:48)
1 = σb1, b(cid:48)
3 = σ4b3. Note that the actual
desorption may occur at force values smaller than the
destabilizing force obtained from linear stability analy-
sis, and thus we introduce the proportionality constant
b(cid:48)
0 in the above expression in order to compare it with
numerical simulations.
The above expression shows how the critical desorption
force F c
z is expected to depend on various properties of
the system, like duty ratio Ωd, bending stiffness κ, and
motor velocity v0. The simulation data for adsorption-
desorption phase diagrams fits well with Eq.(5) [see Figs
(1) and (2) ].
In the limit of v0 = 0 and Ωd following
an equilibrium on-off process due to a sticky surface, the
above expression describes passive desorption of a stiff
filament [35]. In the limit of large v0 and κ, the relation
simplifies to
(cid:48)
0fdNb/
F c
z ≈ b
1 − Ωd
(6)
where Nb denotes the total number of bound motors at
the onset of instability. This expression contains only
one unknown parameter b(cid:48)
0, and thus we shall present
fitting of this expression with simulated data obtained in
the relevant limit. As it turns out, the simulated phase
diagram, in the large v0 and κ limit, is captured well by
the expression obtained above [Fig. 1(c)].
(cid:112)
VI. OUTLOOK
Using mean field theory and linear stability analysis in
one hand, and a stochastic MD simulation in the other,
we investigated the adsorption-desorption transition of
a semiflexible polymer attached to and actively driven
by a bed of molecular motors. We have shown that the
non-equilibrium transition is arguably a continuous tran-
sition. This is characterized by a gradual change in the
fraction of bound motors or desorbed length with increas-
ing pulling force, an absence of phase coexistence, and
increasing correlation time as one approaches the critical
point. We obtained the dependence of desorption force
F c
z as a function of the bending rigidity κ, duty ratio Ωd
and active velocity v0. Phase diagrams obtained from
detailed numerical simulations showed good agreement
with our theory.
The model we studied is closely related to micro-
tubule (MT) organization in animal cells, particularly
those MTs which grow from the microtubule originat-
ing centers (MTOC) towards the cell membrane, and get
captured by the membrane associated dyenein motors.
These motors grab the MT, and tries to walk towards
the MTOC by pulling MTs towards the cell membrane.
Qualitative understanding from our study still remains
valid in such scenarios.
Further, our model may be extended to understand
cell adhesion in presence of elastic relaxation of cell mem-
branes, as opposed to the rigid membranes considered in
the seminal work by Bell [60]. This might be achieved
by considering two semiflexible filaments, as one dimen-
sional projection of two dimensional membranes, and re-
placing the irreversibly attached motor proteins by freely
diffusing reversible bonds.
Our choice of biologically relevant parameter values
makes the current study an interesting prospect for ex-
perimental verification, e.g., in microtubule-kinesin glid-
ing assays. Variation of v0 and Ωd may be achieved by
changing ATP concentration. Bending rigidity κ is par-
tially tunable changing the ambient electrolyte concen-
tration. Work is on to extend our model to study two-
dimensional collective motion of semiflexible filaments
driven by molecular motors. Particular questions as to
how defects in activity of molecular motors [19] impact
motility of single polymers and in turn the collective mo-
tion, will be studied.
Acknowledgments
This work was initiated from a discussion of DC
with Frank Julicher and Debashish Chowdhury of IIT-
Kanpur. We thank Frank Julicher for valuable discus-
sions, and Guillaume Salbreux for useful comments and
a critical reading of the manuscript.
7
FIG. 5:
(Color online) Force-velocity data for kinesin
molecules at 2 mM ATP concentration extracted from
Ref. [12]. The line is a fit to Eq. A3 with v0 = 0.807 µm s−1,
d0 = 0.01 and fs = 1.16 pN.
Appendix A: Non-linear force velocity relation
The dependence of the velocity of procesive kinesin mo-
tor on the ambient ATP concentration was successfully
reproduced by Michaelis-Menten kinetics [12]. This de-
scribes the binding of kinesin motor head M (enzyme)
to an ATP molcule (fuel) and the subsequent ATP-
hydrolysis
M + ATP
ra
(cid:71)(cid:71)(cid:71)(cid:71)(cid:71)(cid:71)(cid:66)(cid:70)(cid:71)(cid:71)(cid:71)(cid:71)(cid:71)(cid:71)
rd
rh
M.ATP
(cid:71)(cid:71)(cid:71)(cid:71)(cid:71)(cid:71)(cid:65)M + ADP + Pi.
Here ra (rd) is rate constant for binding (unbinding) of
an ATP to the kinesin head, and rh is the rate constant
of ATP-hydrolysis; Pi denotes the phosphate ion.
This leads to the Michaelis-Menten expression for the
motor-velocity,
v(f ) = d rh(f )Ψ([AT P ])
(A1)
where d is
the step-size by which the molecular
motor moves per ATP hydrolysis, KM = (rh +
rd)/ra is the Michaelis-Menten constant, Ψ([AT P ]) =
[AT P ]/([AT P ] + KM ) with [AT P ] denoting the ATP
concentration.
The net time scale of ATP hydrolysis has two com-
ponents, one is the time required in absence of force t1,
the other one is an exponentially increased time scale
t2 exp(f δ/kBT ) to cross the enhanced energy barrier f δ
(δ is a characteristic molecular length scale) using ther-
mal energy kBT . Thus the total time per ATP hydrolysis
is t(f ) = t1+t2 exp(f δ/kBT ) with the corresponding rate
rh(f ) = 1/t(f ). This leads to the following general form
of the rate of ATP hydrolysis [12],
rh(f ) =
rh(0)
1 + d0 exp(f /fs)
(A2)
0.10.30.50.70.90246810vM(µm/s)f(pN)kinesinFIG. 6: (Color online) Velocity of the polymer as a function of
external force in one-dimension. Error bars indicate numerical
errors at some representative values of simulation data. The
dashed line is the MFT prediction.
FIG. 7: (Color online) Average number of attached motors as
a function of external force in one-dimension. We show error
bars at representative numerical data. The dashed line shows
the MFT prediction.
where fs = kBT /δ, rh(0) = 1/t1, d0 = t2/t1. The load-
free sliding velocity of a single motor is coupled to ATP-
hydrolysis by v0 = rh(0)d Ψ([AT P ]). Self propulsion v0
is thus a function of ATP concentration, and at high
enough concentration saturates to v0 = rh(0)d. Thus the
motor velocity
vM (f ) =
v0
1 + d0 exp(f /fs)
.
(A3)
Fitting this form with kinesin force-velocity data ob-
tained at large ATP concentration of 2 mM gives v0 =
0.807 µm s−1, d0 = 0.01 and fs = 1.16 pN [12]. The max-
imal force generated by single kinesin molecule is αvM (f )
where α denotes the viscosity of ambient medium. We
used this expression and above-mentioned parameter val-
ues to model active force generation of molecular motors
in the main text.
Note that the above-mentioned chemical reaction de-
scribes the motion of motor head when it is attached to
the polymer track.
In the detached state, it performs
simple diffusion.
Appendix B: 1D characterization
We perform stochastic MD simulations, of an one di-
mensional (1D) rigid rod, under external force Fx, and
activity of the molecular motors. When attached to the
filament, the head of molecular motors walk towards pos-
itive x-direction with active velocity va while the tail
remains irreversibly attached to the substrate, thereby
pushing the filament towards the opposite direction. A
positive Fx thus acts like an opposing load to the mo-
tor driven motion of the filament, whereas a negative Fx
assists that motion. The stochastic noise acts on the fil-
ament as a whole to maintain the ambient temperature.
In the following, we analyze the motion using mean field
theory (MFT) in the over damped limit, and compare
the predictions with simulation results.
The head of each motor protein, when attached to the
filament may move dragged by the filament moving with
velocity v, or due to its active relative motion va. Thus
with respect to the attachment point on the substrate,
the extension of the head position is
dx
dt
= v + va.
(B1)
The resultant force on the filament due to Nb attached
motors is
f a = −kmx Nb.
The filament velocity is given by the force balance
αv = f a + Fx.
(B2)
(B3)
The number of bound motors obey the master equation
∂tNb = ωonN − (ωon + ωoff )Nb,
(B4)
where N is the total number of motor proteins. At steady
state, ∂tNb = 0 implies
Nb = N ωon/(ωon + ωoff ),
(B5)
and dx/dt = 0 implies a negative velocity of the filament
v = −va. On an average, the filament moves in the
direction opposite to the motors with the velocity of a
single free motor protein.
In the presence of external force, the motor detachment
rate ωoff (x) = ω0 exp(kmx/fd), and the active velocity
va(x) = v0/[1 + d0 exp(kmx/fs)]. Using Eq.B2 and v =
−va, Eq.B3 leads to
,
(B6)
kmx =
8
Fx + αva(x)
Nb(x)
−10123−300−100100300−v/v0Fx(pN)simulationMFT5560657075−300−100100300NbFx(pN)simulationMFTwhere Nb can be expressed as a function of x via Eq.B5
and ωoff (x). The non-linear algebraic relation Eq.B6 can
be solved for x. This in turn gives the value of va(x) and
therefore the filament velocity v at a given value of Fx.
This further allows us to calculate Nb as a function of Fx
through Eq. B5.
Using the parameters ωon = 20 s−1 ω0 = 1s−1, α =
1 pN s µm−1, d0 = 0.01, fs = 1.16 pN, fd = 6 pN, and set-
ting the total number of available motor proteins N = 75
we plot the Fx dependence of v and Nb in Figs. 6, and
7 respectively. The plots show comparison of this MFT
prediction with simulation results, and we find reason-
ably good agreement over a broad range of Fx. The
number of attached motors Nb reduces with increase in
the load force, and thus is independent of the sign of
Fx. In absence of external force, the filament moves with
velocity v = −v0 as expected, and the speed reducing
with increasing opposing load Fx > 0. However for as-
sisting load Fx < 0, MFT predicts a v independent of
Fx. Though for Fx > 0 we see good agreement between
MFT and simulations, for large negative force Fx we find
qualitative deviation. At large external forces, the steady
state assumption dx/dt = 0 does not hold, and velocity
of the filament is expected to be v ∼ Fx independent
of the active force. The deviation from MFT shows a
precursor of this crossover.
[1] D. Chowdhury, Physics Reports 529, 1 (2013).
[2] R. D. Vale, Cell 112, 467 (2003).
[3] D. A. Fletcher and R. D. Mullins, Nature 463, 485
(2010).
[4] F. Huber, J. Schnauss, S. Ronicke, P. Rauch, K. Muller,
C. Futterer, and J. Kas, Advances in Physics 62, 1
(2013).
[5] M. L. Gardel, J. H. Shin, F. C. MacKintosh, L. Mahade-
van, P. Matsudaira, and D. A. Weitz, Science 304, 1301
(2004).
[6] F. C. MacKintosh, J. Kas, and P. A. Janmey, Phys. Rev.
Lett. 75, 4425 (1995).
[7] B. L. Goode, D. G. Drubin, and G. Barnes, Current opin-
ion in cell biology 12, 63 (2000).
[8] J. Howard, Mechanics of motor proteins and the cy-
toskeleton (Sinauer Associates, Sunderland, MA, 2001).
[9] K. Gowrishankar, S. Ghosh, S. Saha, C. Rumamol, S.
Mayor, and M. Rao, Cell 149, 1353 (2012).
[10] A. Chaudhuri, B. Bhattacharya, K. Gowrishankar, S.
Mayor, and M. Rao, Proceedings of the National
Academy of Sciences 108, 14825 (2011).
[11] K. Oiwa, S. Chaen, E. Kamitsubo, T. Shimmen, and H.
Sugi, Proceedings of the National Academy of Sciences
87, 7893 (1990).
[12] M. Schnitzer, K. Visscher, and S. Block, Nature cell bi-
ology 2, 718 (2000).
[13] T. Gu´erin, J. Prost, P. Martin, and J.-F. Joanny, Current
opinion in cell biology 22, 14 (2010).
[21] L. Bourdieu, M. Magnasco, D. Winkelmann, and A.
Libchaber, Physical Review E 52, 6573 (1995).
[22] L. Bourdieu, T. Duke, M. Elowitz, D. Winkelmann, S.
Leibler, and A. Libchaber, Physical Review Letters 75,
176 (1995).
[23] E. L. DeBeer, a. M. Sontrop, M. S. Kellermayer, C.
Galambos, and G. H. Pollack, Cell motility and the cy-
toskeleton 38, 341 (1997).
[24] V. Schaller, C. Weber, C. Semmrich, E. Frey, and A. R.
Bausch, Nature 467, 73 (2010).
[25] Y. Sumino, K. H. Nagai, Y. Shitaka, D. Tanaka, K.
Yoshikawa, H. Chat´e, and K. Oiwa, Nature 483, 448
(2012).
[26] H. Yamaoka, S. Matsushita, Y. Shimada, and T. Adachi,
Biomechanics and modeling in mechanobiology 11, 291
(2012).
[27] J. Kierfeld, K. Frentzel, P. Kraikivski, and R. Lipowsky,
The European Physical Journal Special Topics 157, 123
(2008).
[28] P. Kraikivski, R. Lipowsky, and J. Kierfeld, Physical Re-
view Letters 96, 258103 (2006).
[29] S. Banerjee, M. C. Marchetti, and K. Muller-Nedebock,
Physical Review E 84, 011914 (2011).
[30] A. Vilfan, Biophysical journal 97, 1130 (2009).
[31] L. Liu, E. Tuzel, and J. L. Ross, Journal of physics.
Condensed matter : an Institute of Physics journal 23,
374104 (2011).
[32] S. Grill, K. Kruse, and F. Julicher, Physical Review Let-
[14] E. L. F. Holzbaur and Y. E. Goldman, Current opinion
ters 94, 1 (2005).
in cell biology 22, 4 (2010).
[33] X. Oyharcabal and T. Frisch, Physical Review E 71,
[15] M. Badoual, F. Julicher, and J. Prost, Proceedings of
the National Academy of Sciences of the United States
of America 99, 6696 (2002).
[16] A. Vilfan, E. Frey, and F. Schwabl, The European Phys-
ical Journal B 3, 535 (1998).
[17] M. J. I. Muller, S. Klumpp, and R. Lipowsky, Proceed-
ings of the National Academy of Sciences of the United
States of America 105, 4609 (2008).
036611 (2005).
[34] P. Benetatos and E. Frey, Physical review. E, Statistical,
nonlinear, and soft matter physics 67, 051108 (2003).
[35] J. Kierfeld, Physical Review Letters 97, 1 (2006).
[36] C. Friedsam, A. D. C. Bcares, U. Jonas, M. Seitz, and
H. E. Gaub, New Journal of Physics 6, 9 (2004).
[37] S. Cui, C. Liu, and X. Zhang, Nano Letters 3, 245 (2003).
[38] M. Aliee and A. Najafi, Physical Review E 78, 051802
[18] C. Leduc, N. Pavin, F. Julicher, and S. Diez, Physical
(2008).
Review Letters 105, 128103 (2010).
[39] C. Heussinger, M. Bathe, and E. Frey, Phys. Rev. Lett.
[19] L. Scharrel, R. Ma, R. Schneider, F. Julicher, and S. Diez,
99, 048101 (2007).
Biophysical Journal 107, 365 (2014).
[40] T. Nishizaka, R. Seo, H. Tadakuma, K. Kinosita, and S.
[20] S. J. Kron and J. a. Spudich, Proceedings of the National
Academy of Sciences of the United States of America 83,
6272 (1986).
Ishiwata, Biophysical journal 79, 962 (2000).
[41] U. Seifert and R. Lipowsky, Phys. Rev. A 42, 4768
(1990).
9
[42] U. Seifert, Physical Review Letters 84, 2750 (2000).
[43] T. Erdmann and U. S. Schwarz, Physical Review Letters
92, 108102 (2004).
Letters 94, 077804 (2005).
[53] D. Chaudhuri, Physical Review E 75, 21803 (2007).
[54] S. K. Ghosh, K. Singh, and A. Sain, Phys. Rev. E 80,
[44] T. Erdmann and U. S. Schwarz, Journal of Chemical
051904 (2009).
Physics 121, 8997 (2004).
[55] T. B. Liverpool, R. Golestanian, and K. Kremer, Phys.
[45] A.-S. Smith and U. Seifert, Soft Matter 3, 275 (2007).
[46] D. Marenduzzo, S. M. Bhattacharjee, A. Maritan, E. Or-
landini, and F. Seno, Phys. Rev. Lett. 88, 028102 (2001).
[47] R. Kapri, S. M. Bhattacharjee, and F. Seno, Phys. Rev.
Lett. 93, 248102 (2004).
Rev. Lett. 80, 405 (1998).
[56] J. Kierfeld, O. Niamploy, V. Sayakanit, and R. Lipowsky,
The European physical journal. E, Soft matter 14, 17
(2004).
[57] F. Gittes, B. Mickey, J. Nettleton, and J. Howard, J. Cell
[48] S. Kumar and G. Mishra, Phys. Rev. Lett. 110, 258102
Biol. 120, 923 (1993).
(2013).
[58] P. Bak, C. Tang, and K. Wiesenfeld, Phys. Rev. Lett.
[49] A. Dhar and D. Chaudhuri, Phys. Rev. Lett. 89, 65502
59, 381 (1987).
(2002).
[50] P. Ranjith and P. Kumar, Phys. Rev. Lett. 89, 018302
(2002).
[59] P. Bursac, G. Lenormand, B. Fabry, M. Oliver, D. A.
Weitz, V. Viasnoff, J. P. Butler, and J. J. Fredberg, Nat
Mater 4, 557 (2005).
[51] P. Ranjith, P. Kumar, and G. Menon, Phys. Rev. Lett.
[60] G. I. Bell, Science (New York, N.Y.) 200, 618 (1978).
94, 138102 (2005).
[52] O. Hallatschek, E. Frey, and K. Kroy, Physical Review
10
|
1509.06578 | 5 | 1509 | 2016-03-15T01:31:37 | Motility-driven glass and jamming transitions in biological tissues | [
"physics.bio-ph",
"cond-mat.dis-nn",
"cond-mat.soft"
] | Cell motion inside dense tissues governs many biological processes, including embryonic development and cancer metastasis, and recent experiments suggest that these tissues exhibit collective glassy behavior. To make quantitative predictions about glass transitions in tissues, we study a self-propelled Voronoi (SPV) model that simultaneously captures polarized cell motility and multi-body cell-cell interactions in a confluent tissue, where there are no gaps between cells. We demonstrate that the model exhibits a jamming transition from a solid-like state to a fluid-like state that is controlled by three parameters: the single-cell motile speed, the persistence time of single-cell tracks, and a target shape index that characterizes the competition between cell-cell adhesion and cortical tension. In contrast to traditional particulate glasses, we are able to identify an experimentally accessible structural order parameter that specifies the entire jamming surface as a function of model parameters. We demonstrate that a continuum Soft Glassy Rheology model precisely captures this transition in the limit of small persistence times, and explain how it fails in the limit of large persistence times. These results provide a framework for understanding the collective solid-to-liquid transitions that have been observed in embryonic development and cancer progression, which may be associated with Epithelial-to-Mesenchymal transition in these tissues. | physics.bio-ph | physics | Motility-driven glass and jamming transitions in biological tissues
Dapeng Bi1,2, Xingbo Yang1,3, M. Cristina Marchetti1,4, M. Lisa Manning1,4
1Department of Physics, Syracuse University, Syracuse, NY, USA
2Present address: Center for Studies in Physics and Biology, Rockefeller University, NY ,USA
3Present address: Department of Physics, Northwestern University, Evanston, IL
4Syracuse Biomaterials Institute, Syracuse, NY, USA
6
1
0
2
r
a
M
5
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
5
v
8
7
5
6
0
.
9
0
5
1
:
v
i
X
r
a
Cell motion inside dense tissues governs many biological processes, including embryonic development and
cancer metastasis, and recent experiments suggest that these tissues exhibit collective glassy behavior. To make
quantitative predictions about glass transitions in tissues, we study a self-propelled Voronoi (SPV) model that
simultaneously captures polarized cell motility and multi-body cell-cell interactions in a confluent tissue, where
there are no gaps between cells. We demonstrate that the model exhibits a jamming transition from a solid-like
state to a fluid-like state that is controlled by three parameters: the single-cell motile speed, the persistence
time of single-cell tracks, and a target shape index that characterizes the competition between cell-cell adhesion
and cortical tension. In contrast to traditional particulate glasses, we are able to identify an experimentally
accessible structural order parameter that specifies the entire jamming surface as a function of model parameters.
We demonstrate that a continuum Soft Glassy Rheology model precisely captures this transition in the limit of
small persistence times, and explain how it fails in the limit of large persistence times. These results provide
a framework for understanding the collective solid-to-liquid transitions that have been observed in embryonic
development and cancer progression, which may be associated with Epithelial-to-Mesenchymal transition in
these tissues.
Recent experiments have revealed that cells in dense bio-
logical tissues exhibit many of the signatures of glassy ma-
terials, including caging, dynamical heterogeneities and vis-
coelastic behavior [2, 3, 36, 43, 44]. These dense tissues,
where cells are touching one another with minimal spaces in
between, are found in diverse biological processes including
wound healing, embryonic development, and cancer metasta-
sis.
In many of these processes, tissues undergo an Epithelial-
to-Mesenchymal Transition (EMT), where cells in a solid-
like, well-ordered epithelial layer transition to a mesenchy-
mal, migratory phenotype with less well-ordered cell-cell in-
teractions [52, 53], or an inverse process, the Mesenchymal-
to-Epithelial Transition (MET). Over many decades, detailed
cell biology research has uncovered many of the signaling
pathways involved in these transitions [17, 34], which are im-
portant in developing treatments for cancer and congenital dis-
ease.
Most previous work on EMT/MET has focused, however,
on properties and expression levels in single cells or pairs of
cells, leaving open the interesting question of whether there is
a collective aspect to these transitions: Are some features of
EMT/MET generated by large numbers of interacting cells?
Although there is no definitive answer to this question, sev-
eral recent works have suggested that EMT might coincide
with a collective solid-to-liquid jamming transition in biolog-
ical tissues [18, 36, 38, 42]. Therefore, our goal is to develop
a framework for jamming and glass transitions in a minimal
model that accounts for both cell shapes and cell motility, in
order to make predictions that can quantitatively test this con-
jecture.
Jamming occurs in non-biological particulate systems (such
as granular materials, polymers, colloidal suspensions, and
foams) when their packing density is increased above some
critical threshold, and glass transitions occur when the fluid is
cooled below a critical temperature. Over the past 20 years
these phenomena have been unified by "jamming phase dia-
grams" [29, 55].
Building on these successes, researchers have recently used
self-propelled particle (SPP) models to describe dense biolog-
ical tissues [4, 15, 20, 45, 48, 51]. These models are similar
to those for inert particulate matter – cells are represented as
disks or spheres that interact with an isotropic soft repulsive
potential – but unlike Brownian particles in a thermal bath,
self-propelled particles exhibit persistent random walks.
SPP models typically exhibit a glass transition from a dif-
fusive fluid state to an arrested sub-diffusive solid that is con-
trolled by (1) the strength of self-propulsion [15, 20, 35] and
(2) the packing density φ [6, 13, 14, 20, 35]. Just like in ther-
mal systems, a jamming transition occurs at a critical packing
density φG, but this critical density is altered by the persistence
time of the random walks [6, 13, 14, 20, 35].
During many biological processes, however, a tissue re-
mains at confluence (packing fraction equal to unity) while
it changes from a liquid-like to a solid-like state or vice-versa.
For example, in would healing, cells collectively organize to
form a 'moving sheet' without any change in their packing
density [26], and during vertebrate embryogenesis mesendo-
derm tissues are more fluid-like than ectoderm tissues, despite
both having packing fraction equal to unity [44].
Recently, Bi and coworkers [7] have demonstrated that the
well-studied vertex model for 2-D confluent tissues [8, 12, 22,
31, 33, 49] exhibits a rigidity transition in the limit of zero
cell motility. Specifically, the rigidity of the tissue vanishes
at a critical balance between cortical tension and cell-cell ad-
hesion. An important insight is that this transition depends
sensitively on cell shapes, which are well-defined in the ver-
tex model. While promising, vertex models are difficult to
compare to some aspects of experiments because they do not
incorporate cell motility.
In this work, we bridge the gap between the confluent tis-
sue mechanics and cell motility by studying a hybrid between
2
FIG. 1. Analysis of glassy behavior. (A) The mean-squared displacement of cell centers for Dr = 1 and v0 = 0.1 and various values of
p0 (bottom to top: p0 = 3.5,3.65,3.7,3.75,3.8,3.85) show the onset of dynamical arrest as p0 is changed indicating a glass transition. The
dashed lines indicate a slope of 2(ballistic) and 1(diffusive) on log-log plot. (B) The self-intermediate scattering function at the same values of
p0 shown in (A) shows the emergence of caging behavior at the glass transition. (C) The effective self-diffusivity as function of p0 at v0 = 0.1.
At the glass transition De f f becomes nonzero. (D)The cell displacement map in SPV model for a fluid state very close to the glass transition
(p0 = 3.8, v0 = 0.1 and Dr = 1) over a time window t = 104 corresponding to the structural relaxation at which Fs(t) ≈ 1/2.
the vertex model and the SPP model, that we name Self-
Propelled-Voronoi (SPV) model. A similar model was in-
troduced by Li and Sun [28], and cellular Potts models also
bridge this gap [24, 50], although glass transitions have not
been carefully studied in any of these hybrid systems.
I. THE SPV MODEL
While the vertex model describes a confluent tissue as a
polygonal tiling of space where the degrees of freedom are
the vertices of the polygons, the SPV model identifies each
cell only using the center (rrri) of Voronoi cells in a Voronoi
tessellation of space (Dirichlet domains) [27]. The observa-
tion that Voronoi tessellations can describe cellular patterns
in epithelial tissues was first proposed by Honda [21]. For
a tissue containing N-cells, the inter-cellular interactions are
captured by a total energy which is the same as that in the
vertex model. Since the tessellation is completely determined
by the {rrri}, the total tissue mechanical energy can be fully
expressed as E = E({rrri}):
(cid:2)KA(A(ri)− A0)2 + KP(P(ri)− P0)2(cid:3) .
E =
(1)
N
∑
i=1
The term quadratic in cell area A(ri) results from a combi-
nation of cell volume incompressibility and the monolayer's
resistance to height fluctuations [22]. The term involving
the cell perimeter P(ri) originates from active contractility
of the acto-myosin sub-cellular cortex (quadratic in perime-
ter) and effective cell membrane tension due to cell-cell ad-
hesion and cortical tension (both linear in perimeter). This
gives rise to an effective target shape index that is dimension-
less: p0 = P0/√A0. KA and KP are the area and perimeter
moduli, respectively. For the remainder of this manuscript we
assume p0 is homogenous across a tissue, although heteroge-
neous properties are also interesting to consider [59].
In the vertex model [7], a rigidity transition takes place at
a critical value of p0 = p∗0 ≈ 3.81. When p0 < p∗0, cortical
tension dominates over cell-cell adhesion and the energy bar-
riers for local cell rearrangement and motion are finite. The
tissue then behaves as a elastic solid with finite shear mod-
ulus. When p0 > p∗0, cell-cell adhesion dominates and the
energy barriers for local rearrangements vanish, resulting in
zero rigidity and fluid-like behavior. While the energy func-
tional for cell-cell interactions is identical in the vertex and
SPV models, the two are truly distinct: the local minimum en-
ergy states of the vertex model are not guaranteed to be similar
to a Voronoi tessellation of cell centers, although we do find
them to be very similar in practice. Therefore, we are also
interested in whether a rigidity transition in the SPV model
coincides with the rigidity transition of the vertex model.
We define the effective mechanical interaction force ex-
perienced by cell i as FFFi = −∇∇∇iE (see Appendix A for de-
In contrast to particle-based models, FFFi is non-local
tails).
and non-additive: FFFi cannot be expressed as a sum of pairwise
force between cells i and its neighboring cells. Nevertheless,
one can show that momentum is still precisely conserved by
��-���-���-����������������������������������������������h r2(t)it��-���-���-�������p0De↵��-���-���-���������-���-���-���������-���-���-�������ABCD■■■■■■■■■■■■������������������������������Fs(t)increasingp0increasingp0this energy functional in the absence of the additional self-
propulsion forces introduced below.
In addition to FFFi, cells can also move due to self-propelled
motility. Just as in SPP models, we assign a polarity vector
nnni = (cosθi,sinθi) to each cell; along nnni the cell exerts a self-
propulsion force with constant magnitude v0/µ, where µ is
the mobility (the inverse of a frictional drag). Together these
forces control the over-damped equation of motion of the cell
centers rrri
drrri
dt = µFFFi + v0 nnni.
(2)
The polarity is a minimal representation of the front/rear
characterization of a motile cell [50]. While the precise mech-
anism for polarization in cell motility is an area of intense
study, here we model its dynamics as a unit vector that under-
goes random rotational diffusion
∂tθi = ηi(t)
(cid:104)ηi(t)η j(t(cid:48))(cid:105) = 2Drδ(t −t(cid:48))δi j
(3)
where θi is the polarity angle that defines nnni, and ηi(t) is a
white noise process with zero mean and variance 2Dr. The
value of angular noise Dr determines the memory of stochastic
noise in the system, giving rise to a persistence time scale τ =
1/Dr for the polarization vector n. For small Dr (cid:28) 1, the
dynamics of n is more persistent than the dynamics of the
cell position. At large values of Dr, i.e. when 1/Dr becomes
the shortest timescale in the model, Eq. (2) approaches simple
Brownian motion.
The model can be non-dimensionalized by expressing all
lengths in units of √A0 and time in units of 1/(µKAA0). There
are three remaining independent model parameters: the self
propulsion speed v0, the cell shape index p0, and the rota-
tional noise strength Dr. We simulate a confluent tissue un-
der periodic boundary conditions with constant number of
N = 400 cells (no cell divisions or apoptosis) and assume that
the average cell area coincides with the preferred cell area,
i.e. (cid:104)Ai(cid:105) = A0. This approximates a large confluent tissue in
the absence of strong confinement. We numerically simulate
the model using molecular dynamics by performing 105 inte-
gration steps at step size ∆t = 10−1 using Euler's method. A
detailed description of the SPV implementation can be found
in the Appendix Sec. A.
II. CHARACTERIZING GLASSY BEHAVIOR
We first characterize the dynamics of cell motion within the
tissue by analyzing the mean-squared displacement (MSD) of
cell trajectories. In Fig. 1(a), we plot the MSD as function
of time, for tissues at various values of p0 and fixed v0 = 0.1
and Dr = 1. The MSD exhibits ballistic motion ( slope close
to 2 on a log-log plot) at short times, and plateaus at inter-
mediate timescales. The plateau is an indication that cells are
becoming caged by their neighbors. For large values of p0,
the MSD eventually becomes diffusive (slope =1), but as p0 is
decreased, the plateau persists for increasingly longer times.
3
This indicates dynamical arrest due to caging effects and bro-
ken ergodicity, which is a characteristic signature of glassy
dynamics.
(cid:68)
ei(cid:126)k·∆(cid:126)r(t)(cid:69)
Another standard method for quantifying glassy dynamics
is the self-intermediate scattering function [56]: Fs(k,t) =
. Glassy systems possess a broad range of relax-
ation timescales, which show up as a long plateau in Fs(t)
when it is analyzed at a lengthscale q comparable to the near-
est neighbor distance. Fig 1 (b) illustrates precisely this be-
havior in the SPV model, when (cid:126)k = π/r0, where r0 is the
position of the first peak in the pair correlation function. The
average (cid:104)...(cid:105) is taken temporally as well as over angles of (cid:126)k.
Fs(t) also clearly indicates that there is a glass transition as a
function of p0: at high p0 values Fs approaches zero at long
times, indicating that the structure is changing and the tissue
behaves as a viscoelastic liquid. At lower values of p0, Fs re-
mains large at all timescales, indicating that the structure is ar-
rested and the tissue is a glassy solid. Fig 1 (d) demonstrates
that at the structural relaxation time, the cell displacements
show collective behavior across large lengthscales suggesting
strong dynamical heterogeneity. This is strongly reminiscent
of the 'swirl' like collective motion seen in experiment in ep-
ithelial monolayers [2, 3, 15, 39, 40].
A. A dynamical order parameter for the glass transition
Although the phase space for this model is three dimen-
sional, we now study the model at a fixed value of Dr = 1.
We then search for a dynamical order parameter that dis-
tinguishes between the glassy and fluid states as a func-
tion of the two remaining model parameters,(v0, p0). A
candidate order parameter is the self-diffusivity Ds: Ds =
2(cid:105)/(4t). For practicality, we calculate Ds using
limt→∞(cid:104)∆r(t)
simulation runs of 105 time steps, chosen to be much longer
than the typical caging timescale in the fluid regime. We
present the self-diffusivity in units of D0 = v2
0/(2Dr), which is
the free diffusion constant of an isolated cell. De f f = Ds/D0
then serves as an accurate dynamical order parameter that dis-
tinguishes a fluid state from a solid (glassy) state in the space
of (v0, p0), matching the regimes identified using the MSD
and Fq. In Fig. 2, the fluid region is characterized by a finite
value of De f f and De f f drops below a noise floor of ∼ 10−3
In practice, we label
as the glass transition is approached.
materials with De f f > 10−3 as fluids indicated by an orange
dot, and those with De f f ≤ 10−3 as solids indicated by blue
squares. Importantly, we find that the SPV model in the limit
of zero cell motility shares a rigidity transition with the vertex
model [7] at p0 ≈ 3.81, and that this rigidity transition con-
trols a line of glass transitions at finite cell motilities. Typical
cell tracks (Fig. 2) clearly show caging behavior in the glassy
solid phase.
4
(A) Glassy phase diagram for confluent tissues as function of cell motility v0 and target shape index p0 at fixed Dr = 1. Blue data
FIG. 2.
points correspond to solid-like tissue with vanishing De f f ; orange points correspond to flowing tissues (finite De f f ). The dynamical glass
transition boundary also coincides with the locations in phase space where the structural order parameter q = (cid:104)p/√a(cid:105) = 3.81 (dashed line).
In the solid phase, q ≈ 3.81 and q > 3.81 in the fluid phase. (B) Instantaneous tissue snapshots show the difference in cell shape across the
transition. Cell tracks also show dynamical arrest due to caging in the solid phase and diffusion in the fluid phase.
B. Cell shape is a structural order parameter for the glass
transition
In glassy systems it can be difficult to experimentally dis-
tinguish between a truly dynamically arrested state and a state
with relaxation times longer than the experimental time win-
dow. Similarly, in tissues it is experimentally challenging to
quantify a glass transition through the measurement of a dy-
namical quantity such as the diffusivity Ds. Identifying a static
quantity that directly probes the mechanical properties of a tis-
sue would therefore be a powerful tool for experiments. Pu-
liafito et al. have suggested that shape changes accompany
dynamical arrest in proliferating tissues [41]. Similarly, a
structural signature based on cell shapes – the shape index
q = (cid:104)p/√a(cid:105) – was previously shown to be an excellent or-
der parameter for the confluent tissue rigidity transition in the
vertex model [38]. In a model where cells were not motile
(v0 = 0) we found that when p0 < 3.813, q is constant ∼ 3.81
and when p0 > 3.81 q grows linearly with p0. Quite sur-
prisingly, we found that the prediction of q = 3.813 works
perfectly in identifying a jamming transition in in-vitro ex-
periments involving primary human tissues, where cells are
clearly motile (v0 (cid:54)= 0) [38]. At that time, we did not under-
stand why the v0 = 0 theory worked so well for these tissues.
The prediction of a solid-liquid transition in the SPV
model presented here provides an explanation for this obser-
vation.We find that q (which can be easily calculated in ex-
periments or simulations from a snapshot) can be used as a
structural order parameter for the glass transition for all val-
ues of v0, not just at v0 = 0. Specifically, the boundary defined
by q = 3.813, shown by the blue dashed line in Fig. 2(A) co-
incides extremely well with the glass transition line obtained
using the dynamical order parameter, shown by the round and
square data points. The insets to Fig. 2 also illustrate typical
cell shapes: cells are isotropic on average in the solid phase
and anisotropic in the fluid phase. This highlights the fact that
q can be used as a structural order parameter for the glass tran-
sition at all cell motilities, providing a powerful new tool for
analyzing tissue mechanics.
III. A THREE-DIMENSIONAL JAMMING PHASE
DIAGRAM FOR TISSUES
Having studied the glass transition as function of v0 and
p0 at a large value of Dr, we next investigate the full three-
dimensional phase diagram by characterizing the effect of Dr
on tissue mechanics and structure. Dr controls the persistence
time τ = 1/Dr and persistence length or P´eclet number Pe ∼
v0/Dr of cell trajectories; smaller values of Dr correspond to
more persistent motion.
In Fig. 3(A), we show several 2D slices of the three di-
mensional jamming boundary. Solid lines illustrate the phase
transition line identified by the structural order parameter
q = 3.813 as function of v0 and p0 for a large range of Dr val-
ues (from 10−2 to 103). (In Appendix B 2 we demonstrate that
the structural transition line q = 3.813 matches the dynamical
transition line for all studied values of Dr.) In contrast to re-
sults for particulate matter [6], this figure illustrates that the
glass transition lines meet at a single point (p0 = 3.81) in the
limit of vanishing cell motility, regardless of persistence.
Fig. 3(B) shows an orthogonal set of slices of the jamming
AB■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●����������������������������������FluidSolid5
FIG. 3. (A) The glass transition in v0 − p0 phase space shifts as the persistence time changes. Lines represent the glass transition identified by
the structural order parameter q = 3.81. The phase boundary collapse to a single point at p∗0 = 3.81, regardless of Dr, in the limit v0 → 0. (B)
The glass transition in p0 − Dr phase space shifts as a function of v0 (from top to bottom: v0 = 0.02,0.08,0.14,0.2,0.26) For large v0 there is
a crossover in the behavior at Dr ∼ µKAA0 = 1, as discussed in the main text. (C) The phase boundary between solid and fluid as function of
motility v0, persistence 1/Dr and p0 which is tuned by cell-cell adhesion can be organized into a schematic 3D phase diagram. Red lines on
the surface correspond to iso-v0 contours and blue lines correspond to iso-Dr contours.
diagram, illustrating how the phase boundary shifts as func-
tion of p0 and Dr at various values of v0. This highlights the
interesting result that a solid-like material at high value of Dr
can be made to flow simply by lowering its value of Dr. The
crossover in behavior at large v0 occurs when the persistence
time 1/Dr is approximately equal to the viscous relaxation
time 1/(µKAA0) = 1.
These slices can be combined to generate a three-
dimensional jamming phase diagram for confluent biological
tissues, shown in Fig. 3(C). This diagram provides a concrete,
quantifiable prediction for how macroscopic tissue mechanics
depends on single-cell properties such as motile force, persis-
tence, and the interfacial tension generated by adhesion and
cortical tension.
We note that Fig. 3(C) is significantly different from the
jamming phase diagram conjectured by Sadati et al [42],
which was informed by results from adhesive particulate mat-
ter [55]. For example, in particulate matter adhesion enhances
solidification, while in confluent models adhesion increases
cell perimeters/surface area and enhances fluidization. In ad-
dition, we identify "persistence" as a new axis with a poten-
tially significant impact on cell migration rates in dense tis-
sues.
To better understand why persistence is so important in
dense tissues, we first have to characterize the transitions be-
tween different cellular structures.
In the limit of zero cell
motility, the system can be described by a potential energy
landscape where each allowable arrangement of cell neigh-
bors corresponds to a metastable minimum in the landscape.
There are many possible pathways out of each metastable
state: some of correspond to localized cell rearrangements,
while others correspond to large-scale collective modes. The
maximum energy required to transition out of a metastable
state along each pathway is called an energy barrier [8].
We observe that tissue fluidity can increase drastically with
increasing Dr at finite cell speeds. This suggests that different
(A-C) Instantaneous cell displacements at p0 = 3.65
FIG. 4.
and v0 = 0.5. They are different from the displacements shown in
Fig. 1(D) which are averaged over the structural relaxation timescale.
(A) At the lowest value of Dr = 0.01, the cells are able to flow by
collectively displacing along the 'soft' modes of the system (Ap-
(B) At Dr = 0.1, collective displacements are less
pendix. B 1).
pronounced. (C) For Dr = 1 and larger, the displacements appear
disordered and uncorrelated.
pathways (with lower energy barriers) must become dynami-
cally accessible at higher values of Dr.
One hint about these pathways comes from the instanta-
neous cell displacements, shown for different values of Dr in
Fig. 4. At high values of Dr, (p0 = 3.78, v0 = 0.1) the instan-
taneous displacement field is essentially random and largely
uncorrelated, as shown in Fig. 4, and the material is solid-
like. There is no collective behavior among cells, and each
cell 'rattles' independently near its equilibrium position.
However, as Dr is lowered, the instantaneous displacement
field becomes much more collective (Fig. 4) and the tissue
begins to flow, presumably because these collective displace-
ment fields correspond to pathways with lower transition en-
ergies.
Two obvious questions remain: How does a lower value
of Dr generate more collective instantaneous displacements?
Why should collective instantaneous displacements generi-
Drp0v0���������������������������������������0.01125201000●●●●●●●●●●■■■■■■■■■■◆◆◆◆◆◆◆◆◆◆▲▲▲▲▲▲▲▲▲▲▼▼▼▼▼▼▼▼▼▼○○○○○○○○○□□□□□□□□□□��-���-����������������������������v0p0DrABCp0v01/DrSolid-likeDr=0.0110.1ABCcally have lower energy barriers? The first question can be
answered by extending ideas first proposed by Henkes, Fily
and Marchetti [20] to explain why motion in self-propelled
particle models seems to follow the 'soft modes' of a solid.
This argument is based on a simple, yet powerful observation:
in the limit of zero motility (v0 = 0), a solid-like state will have
a well-defined set of normal modes of vibration (with frequen-
cies {ων}), and a corresponding set of eigenvectors ({ eν}) that
forms a complete basis. At higher motilities (v0 > 0) near the
glass transition, the motion of particles in the system can be
expanded in terms of the eigenvectors. As discussed in Ap-
pendix B 1, one can use this observation to show that in the
limit of Dr → 0, motion along the lowest frequency eigen-
modes is amplified – the amplitude along each mode is pro-
portional to 1/ω2
ν. These low-frequency normal modes are
precisely the collective displacements observed for low Dr.
The second question is more difficult to answer because it
is impossible to enumerate all of the possible transition path-
ways and energy barriers in a disordered material. However,
a partial answer comes from recent work in disordered par-
ticulate matter showing that low-frequency normal modes do
have significantly lower energy barriers [30, 61] than higher
frequency normal modes. A similar analysis could potentially
be performed in vertex or SPV models.
IV. A CONTINUUM MODEL FOR GLASS TRANSITIONS
IN TISSUES
Although continuum hydrodynamic equations of motion
have been developed by coarse graining SPP models in the
dilute limit, there is no existing continuum model for a dense
active matter system near a glass transition. Here we propose
that a simple trap [32] or Soft Glassy Rheology (SGR) [47]
model provides an excellent continuum approximation for the
phase behavior in the large Dr Brownian regime, but fails in
the small Dr limit.
For large Dr it is known that particle behave like Brownian
particles with an effective temperature Te f f = v2
0/2µDr [14].
This mapping becomes exact when Dr → ∞ at fixed "effec-
tive inertia" (µDr)−1 [13].
In other words, like in granular
systems [1, 9], the effective temperature in SPP is dominated
by kinetic effects. Guided by this result we conjecture that
in our model the temperature also scale quadratically with the
velocity,
Te f f ∝ cv2
0.
(4)
Physically, this effective temperature gives the amount of en-
ergy available for individual cells to vibrate within their cage
or 'trap'.
The next important question is how to characterize the 'trap
depths', or energy barriers between metastable states. In the
Brownian regime (large Dr) there is no dynamical mechanism
for the cells to organize collectively, and therefore a reason-
able assumption is that the rearrangements are small and lo-
calized.
In [8], some of us explicitly calculated the statistics of en-
ergy barriers for localized rearrangements in the equilibrium
6
FIG. 5. Comparison between SPV glass transition and an analytic
prediction based on a Soft Glass Rheology (SGR) continuum model.
The dashed line corresponds to an SGR prediction with no fit param-
eter based on previously measured vertex model trap depths [8]. Data
points correspond to SPV simulations with Dr = 10−3 and where we
have defined Te f f = cv2
0 with c = 0.1 as the best-fit normalization pa-
rameter. Blue points correspond to simulations which are solid-like,
with De f f < 10−3, and the boundary of these points define the ob-
served SPV glass transition line. (Inset) L2 difference between SPV
glass transition line (at best-fit value of c) and the predicted SGR
transition line at various values of Dr. The SGR prediction based on
localized T1 trap depths works well in the high Dr limit, but not in
the low Dr limit.
vertex model.
In the 2D vertex model, one can show that
localized rearrangements must occur via so-called T1 tran-
sitions [58]. Using a trap model [32] or Soft Glassy Rheol-
ogy(SGR) [47] framework, we were able to use these statis-
tics to generate an analytic prediction, with no fit parameters,
for the glass transition temperature Tg as function of p0.
To see if the SGR prediction for the glass transition holds
for the SPV model in the large Dr limit, we simply overlay
the data points corresponding to glassy states from the SPV
model with the glass transition Tg line predicted in [8]. There
is one fitting parameter c that characterizes the proportion-
ality constant in Eq. 4. Fig 5, shows that the SPV data for
Dr = 103 is in excellent agreement with our previous SGR
prediction. Because Te f f ∼ v2
0, and the glass transition line
scales as Tg ∼ p∗0 − p0 for p0 (cid:28) p∗0 and Dr → 0, the glass
transition line scales as v0 ∼ (p∗0 − p0)
The reason the effective temperature SGR model works
here is that, like in SPP models of spherical active Brownian
colloids, the angular dynamics of each cell evolves indepen-
dently of cell-cell interactions and of the angular dynamics of
other cells. An additional alignment interaction that couples
the angular and translational dynamics may therefore modify
this behavior.
0.5 in those limits.
To our knowledge, this is the first time that a SGR / trap
model prediction has been precisely tested in any glassy sys-
tem. This is because, unlike most glass models, we can enu-
merate all of the trap depths for localized transition paths in
the vertex model.
■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■�����������������������p0■■■■■■■�������������������DrTeffL2However, for small values of Dr, we have shown that cell
displacements are dominated by collective normal modes, and
therefore the energy barriers for localized T1 transitions are
probably irrelevant in this regime. The inset to Fig 5 shows
the deviation (L2-norm) between glass transition lines in the
SPV model and T1-based SGR prediction as a function of Dr.
We see that the SGR prediction fails in the small Dr limit,
as expected. A better understanding of the energy barriers
associated with collective modes will be required to modify
the theory at small Dr.
V. DISCUSSION AND CONCLUSIONS
We have shown that a minimal model for confluent tissues
with cell motility exhibits glassy dynamics at constant den-
sity. This model allows us to make a quantitative prediction
for how the fluid-to-solid/jamming transition in biological tis-
sues depends on parameters such as the cell motile speed, the
persistence time associated with directed cell motion, and the
mechanical properties of the cell (governed by adhesion and
cortical tension). We define a simple, experimentally acces-
sible structural order parameter – the cell shape index – that
robustly identifies the jamming transition, and we show that a
simple analytic model based on localized T1 rearrangements
precisely predicts the jamming transition in the large Dr limit.
We also show that this prediction fails in the small Dr limit,
because the instantaneous particle displacements are domi-
nated by collective normal modes.
This model makes several experimentally-verifiable predic-
tions for cell shape and tissue mechanics:
• The order parameter q = 3.81 is a structural signature
for the glass transition, even in tissues with significant
cell motility or dynamics. This prediction has already
been tested in epithelial lung tissue [38], but it should
be much more broadly applicable. We have performed
a rudimentary shape analysis of a small number of im-
ages from other systems that have been previously pub-
lished, including proliferating MCDK monolayers [10]
and convergent extension in fruit fly development [11]
and found that the shapes are consistent with this pre-
diction. A much more careful analysis with full data
sets should be performed to further validate this predic-
tion or understand where it breaks down.
• In the limit of vanishing cell motility, shape and pres-
sure fluctuations should vanish when the jamming tran-
sition is approached from the solid side, and remain
zero in the fluid. A finite motility v0 will induce such
fluctuations in the fluid phase, as confirmed by prelimi-
nary calculation of cellular stresses and pressure in the
SPV model [62]. This could be studied by combin-
ing measurement of cell shape fluctuations with traction
force microscopy (TFM) in wound healing assays. Af-
ter locating the glass transition by imaging cell shape
changes, it may be possible to extract information on
cell motility v0 from cellular stresses and pressure in-
ferred from TFM in the fluid phase near the glass tran-
7
sition. This suggests that one may estimate cell motility
by examining the changes in cellular stresses and pres-
sure in the cell monolayer near the unjamming transi-
tion and assuming that the local velocity of the mono-
layer is very small just above the transition. The latter
assumption can also be verified independently via par-
ticle image velocimetry (PIV).
• Cell proliferation, so far neglected in our model, causes
an increase in cell number density in confluent tissues.
Often this is accompanied by a reduction in individual
cell motility v0, via contact inhibition of locomotion. In
cases where this is the dominant effect and changes to
the ratio between A0 and P0 are negligible, our work
predicts that proliferation would drive the system to-
wards jamming. This is consistent with existing reports
in the literature [10], although more work is required to
test the prediction carefully. In tissues where v0 remains
low at all times [41], our model predicts that prolifera-
tion can either cause jamming or unjamming, depend-
ing on whether cell divisions are oriented in such a way
to decrease or increase cell shape anisotropy.
• Spatial correlations and fluctuations of the cell dis-
placement field, such as swirl sizes [2, 3, 15, 39, 40]
, should grow as a tissue approaches the glass transi-
tion from the fluid side. Very recently, a similar pre-
diction for displacements and correlation lengths based
on a particle-based model has been verified in one cell
type [15]. The SPV model, which makes predictions for
cell shapes in addition to displacements and correlation
lengths, could be tested simply by compiling detailed
statistics about cell shapes and cell motion in epithelial
monolayers.
Although all of the work presented here focuses on the SPV
model, which tracks cell centers and therefore has only two
degrees of freedom per cell, we found that in the limit of zero
cell motility it exhibits the same rigidity transition as the ver-
tex model which has two degrees of freedom per vertex. We
have also checked that an "active vertex model", where ac-
tive motile forces are added to the vertex model vertices, also
exhibits a robust glass transition characterized by the shape
order parameter q. The fact that two models with ostensibly
different degrees of freedom share the same transition sug-
gests that there is a deeper universality, perhaps generated by
isostaticity, that remains to be understood.
Another result of this work is the surprising and unexpected
differences between confluent models (such as the vertex and
SPV models) and particle-based models (such as Lennard-
Jones glasses and SPP models). For example, works by
Berthier [6], Fily and Marchetti [14] in SPP models suggest
that the location of the zero motility glass transition packing
density φG (defined as the density at which dynamics cease
in the limit of v0 → 0) depends the value of noise, Dr. This
is also related to the observation that the jamming and glass
transition are not controlled by the same critical point in non-
active systems [23, 37]. We find this is not the case in the SPV
model. Fig. 3(a & b) show that while the glass transition point
p∗0 shifts with Dr at finite values of v0, in the limit of vanishing
motility, all glass transition lines merge on to a single point in
the limit v0 → 0, namely p∗0 = 3.81.
Given these differences, it is important to ask which type
of model is appropriate for a given system. We argue that
SPV models are maybe more appropriate for many biologi-
cal tissues. Whereas SPP models interact with two-body in-
teractions that only depend on particle center positions, both
SPV and vertex models naturally incorporate contractility as
a key property of living cells and capture the inherently multi-
body nature of intercellular forces due to shape deformations.
Unlike equilibrium vertex models, SPV models account for
cell motility, and they are also much easier to simulate in 3D
(which is nearly impossible in practice for the vertex model.)
Recent work by Li and Sun [28] also models a confluent
cell as a Voronoi tessellation of the plane. An important dif-
ference between this work and ours is that in Ref. [28] cell-
cell adhesion is captured via a potential that is quadratic in the
distance between cell centers, just as in particle models. We
might guess that stronger cell-cell adhesion in their model will
result in stiffening of the tissue, which is common for parti-
cle based models, although that remains to be tested in active
systems. In contrast, adhesion enters our model through the
coupling of the shape energy to the cell perimeter. Increas-
ing cell-cell adhesion (or decreasing cortical tension) yields a
larger value of p0, which leads to the tissue becoming softer.
We expect that other shape-based models of confluent tis-
sue dynamics will also yield the glass transition described
here. For example, it has been reported in recent works based
on the Cellular Potts model [24, 50] and in a modified SPP
model [15] when the cell motile force is decreased beyond a
certain threshold, the motion of cells transitions from diffusive
to sub-diffusive. This is similar to crossing the glass transition
line in the SPV model by decreasing the value of v0.
In this work and in previous work based on the vertex
model, the cell volume is generally assumed to be fixed.
While this is a good assumption in developmental systems
such as drosopholia [16, 25] and zebrafish [31], epithelial
tissues cells can show significant volume fluctuations, as re-
ported recently [63, 64]. Therefore, it will be important to
incorporate volume fluctuations in future iterations of the ver-
tex model or the SPV model, as they introduce another source
of active shape fluctuation and could therefore lead to jam-
ming or unjamming of the tissue locally and potentially shift
the location of the rigidity and glass transitions.
In our version of the SPV model, we have assumed that cell
polarity is controlled by simple rotational white noise. It is
also possible to include more complex mechanisms. For ex-
ample, external chemical or mechanical cues could be mod-
eled by coupling v0 and nnni to chemoattractant or mechanical
gradients, allowing waves or other pattern formation mecha-
nisms to interact with the jamming transition. Similarly, sim-
ple alignment rules (such as those in the Viscek model [57])
could lead to collective flocking modes that also affect glassy
dynamics.
Another interesting extension of the SPV model would be
to study the role of cell-cell friction, which has already been
shown to be important in controlling collective dynamics in
8
particle-based tissue models [15]. Our current model includes
viscous frictional coupling of cell to the 2D substrate and cell-
cell adhesion enters as a negative line tension on interfaces.
However, it would be possible to add a frictional force be-
tween cells proportional to the length of the edge shared be-
tween two cells, and we know from previous work on par-
ticulate glasses that these localized frictions can change the
location of jamming/glass transition and the nature of spatial
correlations in a glass [19, 46].
It is also tempting to speculate about the relationship be-
tween the unjamming transition captured by our model and the
epithelial-mesenchimal transition (EMT) that precedes cell
escape from a solid tumor mass. The EMT involves signifi-
cant changes in cell-cell adhesion and cytoskeletal composi-
tion, with associated changes in cell shape and motility. This
suggests that escape from the tumor mass is controlled not
just by the chemical breakdown of the basement membrane,
but also by specific changes in mechanical properties of both
individual cells and the surrounding tissue [60]. One could
then hypothesize that the collective unjamming described here
may provide the first necessary step towards the mechanical
changes needed for cell escape from primary tumors.
In particular, recent work suggests that cancer tumors are
mechanically heterogeneous, with mixtures of stiff and soft
cells that have varying degrees of active contractility [59]. Our
jamming phase diagram suggests that the soft cells, which of-
ten exhibit mesenchymal markers and presumably correspond
to higher values of p0, might unjam and move towards the
boundary of a primary tumor more easily than their stiff coun-
terparts. Examining the effects of tissue heterogeneity on tis-
sue rigidity and patterns of cell motility is therefore a very
promising avenue for developing predictive theories for tumor
invasiveness and metastasis.
Appendix A: Simulation algorithm for the SPV model
To create an initial configuration for the simulation, we first
generate a seed point pattern using random sequential addi-
tion (RSA) [54] and anneal it by integrating Eq. 2 with v0 = 0
for 100 MD steps. The resulting structure then serves as an
initially state for all simulations runs. The use of (RSA) only
serves to speed up the initial seed generation as using a Pois-
son random point pattern does not change the results presented
in this paper.
At each time step of the simulation, a Voronoi tesselation is
created based on the cell centers. The intercellular forces are
then calculated based on shapes and topologies of the Voronoi
cells (see discussion below). We employ Euler's method to
carry out the numerical integration of Eq. 2, i.e., at each time
step of the simulation the intercellular forces is calculated
based on the cell center positions in the previous time step.
In a Delaunay triangulation, a trio of neighboring Voronoi
centers define a vertex of a Voronoi polygon. For example in
Fig. 6, ((cid:126)ri,(cid:126)r j,(cid:126)rk) define the vertex(cid:126)h3, which is given by
(cid:126)h3 = α(cid:126)ri + β(cid:126)r j + γ(cid:126)rk,
(A1)
where the coefficients are given by
α = (cid:107)(cid:126)r j −(cid:126)rk(cid:107)2((cid:126)ri −(cid:126)r j)· ((cid:126)ri −(cid:126)rk)/D
β = (cid:107)(cid:126)ri −(cid:126)rk(cid:107)2((cid:126)r j −(cid:126)ri)· ((cid:126)r j −(cid:126)rk)/D
γ = (cid:107)(cid:126)ri −(cid:126)r j(cid:107)2((cid:126)rk −(cid:126)ri)· ((cid:126)rk −(cid:126)r j)/D
D = 2(cid:107)((cid:126)ri −(cid:126)r j)× ((cid:126)r j −(cid:126)rk)(cid:107)2
.
(A2)
9
since the interaction is inherently multi-cellular in nature and
interactions between i and j also depend on k and l (see
Fig. 6).
For the typical configuration shown in Fig. 6, the first term
in Eq. A5 can be expanded using the chain rule and calculated
using Eq. A1
∂E j
∂riµ
= ∑
ν
∂h2ν
∂riµ
+
∂E j
∂h3ν
∂h3ν
∂riµ
.
(A7)
(cid:18) ∂E j
∂h2ν
(cid:19)
In Eq. A7, only terms involving (cid:126)h2 and (cid:126)h3 are kept since E j
does not depend on other vertices of cell i. ν is a cartesian
coordinate index. The energy derivative in Eq. A7 can be cal-
culated in a straightforward way, by using Eqs. A3 and A4
FIG. 6. Cell centers positions are specified by vectors {(cid:126)r}. They
form a Delaunay triangulation (black lines). Its dual is the Voronoi
tessellation (red lines), with vertices given by {(cid:126)h}.
∂E j
∂h2x
and
∂E j
∂h2y
In the vertex model, the total mechanical energy of a tissue
depends only on the areas and perimeters of cells:
(cid:2)KP(Ai − A0)2 + KP(Pi − P0)2(cid:3) .
E =
N
∑
i=1
(A3)
In a Voronoi tessellation, the area and perimeter of a cell i can
be calculated in terms of the vertex positions
+ 2KP(Pj − P0)
∂Pj
∂h2x
∂A j
∂h2x
=2KA(A j − A0)
=KA(A j − A0)(h3y − h7y)
+ 2KP(Pj − P0)
(cid:32)
(cid:33) (A8)
h2x − h7x
(cid:107)(cid:126)h7 −(cid:126)h2(cid:107)
+
h2x − h3x
(cid:107)(cid:126)h2 −(cid:126)h3(cid:107)
+ 2KP(Pj − P0)
∂Pj
∂h2y
h2y − h7y
(cid:107)(cid:126)h7 −(cid:126)h2(cid:107)
+
h2y − h3y
(cid:107)(cid:126)h2 −(cid:126)h3(cid:107)
(cid:33)
.
(A9)
∂A j
∂h2y
=2KA(A j − A0)
=KA(A j − A0)(h3x − h7x)
+ 2KP(Pj − P0)
(cid:32)
Similarly, the second term on the r.h.s. of Eq. A5 can be cal-
culated in a similar way.
Pi =
Ai =
zi−1
∑
m=0(cid:107)(cid:126)hm −(cid:126)hm+1(cid:107);
zi−1
1
∑
m=0(cid:107)(cid:126)hm ×(cid:126)hm+1(cid:107),
2
(A4)
Appendix B: Cell displacements and structural order parameter
as a function of Dr
1. Expanding cell displacements in an eigenbasis associated
with the underlying dynamical matrix
where zi is the number of vertices for cell i (also number of
neighboring cells) and m indexes the vertices. We use the con-
vention(cid:126)hzi =(cid:126)h0.
With these definitions, the total force on cell-i can be cal-
culated using Eq. A3
Fiµ ≡ −
∂E
∂riµ
= − ∑
j∈n.n.(i)
∂E j
∂riµ −
∂Ei
∂riµ
,
(A5)
here µ denotes the cartesian coordinates (x,y). The first term
on the r.h.s. of Eq. A5 sums over all nearest neighbors of cell
i. It is the force on cell i due to changes in neighboring cell
shapes. The second term is the force on cell i due to shape
changes brought on by its own motion.
It maybe tempting to treat ∂E j
∂riµ
as the force between cells-i
and j, but
∂E j
∂riµ (cid:54)=
∂Ei
∂r jµ
(A6)
In the absence of activity (v0 = 0), the tissue is a solid
for p0 < p∗0 = 3.81. As v0 is increased, the solid behavior
persists up to v0 = v∗0(p0), which is given by the glass tran-
sition line in Fig. 2.
In order for the tissue to flow, suffi-
cient energy input is needed to overcome energy barriers in
the potential energy landscape, which are a property of the
underlaying solid state at v0 = 0.
In this limit, the instan-
taneous cell center positions {(cid:126)ri(t)} can be thought of as a
small displacement {(cid:126)di(t)} from the nearest solid reference
state {(cid:126)r0i} [20] where (cid:126)di(t) = (cid:126)ri −(cid:126)r0i. The (cid:126)r0i correspond
to positions of cell in a solid, which has a well-defined lin-
ear response regime [7]. The linear response is most con-
veniently expressed as the eigen-spectrum of the dynamical
matrix Di jαβ. Since the eigenvectors { ei,ν} of Di jαβ form a
complete orthonormal basis, the cell center displacement can
then be expressed as a linear combination of { ei,ν}
(cid:126)di(t) = ∑
ν
aν(t) ei,ν
(B1)
~ri~rj~rk~rl~h1~h2~h3~h4~h5~h6~h7~h8to get the ensemble averaged solution for the amplitude be-
comes
10
(cid:104)aν(t)(cid:105) = aν(0)e−kt + v0 cos (θν − ψ(0))
In the limit of Dr → ∞ and Eq. B11 becomes
e−kt − e−Drt
Dr − k
.
(B11)
(cid:104)aν(t)(cid:105) = aν(0)e−kt .
(B12)
νt +
v0
µω2
ν
cos (θν − ψ(0))
This suggests that while normal modes control the rate of de-
cay, they do no affect the long-time behavior.
However as Dr → 0, Eq. B11 becomes
aν(t) = aν(0)e−µω2
νt(cid:17)
(cid:16)
1− e−µω2
(B13)
The second term in this equation scales as ∼ 1/ω2
ν. Therefore,
at short times (corresponding to instantaneous response), the
mode amplitude aν is much larger for modes at lower frequen-
cies. Since the reference state is an elastic solid with Debye
scaling D(ω) ∼ ω as ω → 0 [7], this suggests that the dis-
placement will be heavily dominated by the lowest frequency
modes that are spatially more collective in nature.
For simplicity, we will adopt the Bra-ket notation and express
the eigenbasis simply as ν(cid:105) and Eq. B1 becomes
where
d(cid:105) = ∑
ν
aν(t)ν(cid:105),
Dν(cid:105) = ω2
νν(cid:105)
(B2)
(B3)
and ω2
ν are the eigenvalues of the dynamical matrix.
The polarization vector ni can also be expressed as a linear
combination of eigenvectors
n(cid:105) = ∑
ν
bν(t)ν(cid:105).
(B4)
Since the polarization vector and eigenvector are both unit
vectors, it follows that bν(t) = (cid:104)nν(cid:105) = cos(θν − ψ). Where ψ
is the angle of the polarization and θν the angle of the eigen-
vector.
Then the equation of motion for (cid:126)di (Eq. 2), can be rewritten
as
(cid:126)d = −µ
∂E
∂(cid:126)ri
Using Eqs. B2–B5, we find
(cid:12)(cid:12)(cid:12)(cid:12)(cid:126)r0,i
+ v0 ni
(B5)
2. Effect of Dr on glass transition boundary
d
dt (cid:104)νd(cid:105) = −µ(cid:104)ν D d(cid:105) + v0(cid:104)νn(cid:105) ,or
d
dt
aν(t) = −µω2
νaν(t) + v0bν(t).
Then the equation of motion for each amplitude is
d
dt
aν(t) = −µω2
νaν(t) + v0cos(θν − ψ)
ψ = η.
(B6)
(B7)
FIG. 7. Comparison between glass transition boundaries obtained
using shape order parameter (red line) and De f f (blue squares and
orange circles).
This is just the equation of motion for a self-propelled particle
tethered to a spring with active forcing that is strongest along
the direction of the eigenvector [5]. The solution is then:
Figure. 7 shows that the location in phase space where the
shape index q = 3.81 is in excellent agreement with the dy-
namical solid-fluid phase boundary determined by De f f , at all
values of Dr.
aν(t) = aν(t = 0)e−kt + v0
dt(cid:48)e−k(t−t(cid:48))cos(θν − ψ), (B8)
ACKNOWLEDGMENTS
(cid:90) t
0
(cid:90) t
0
where k = µω2
ν.
Solving for the ensemble averaged quantity:
(cid:104)aν(t)(cid:105) = aν(t = 0)e−kt + v0
and using the relations
dt(cid:48)e−k(t−t(cid:48))(cid:104)cos(θν − ψ)(cid:105),
(B9)
(cid:104)cosψ(t)(cid:105) = cosψ(0)e−Drt;
(cid:104)sinψ(t)(cid:105) = sinψ(0)e−Drt
cos(θν − ψ) = sin(θν)sin(ψ) + cos(θν)cos(ψ),
(B10)
M.L.M. acknowledges support from the Alfred P. Sloan
Foundation. M.L.M and D.B. acknowledge support from
NSF-BMMB-1334611 and NSF-DMR-1352184, the Gordon
and Betty Moore Foundation and the Research Corpora-
tion for Scientific Advancement and NIH R01GM117598-
02. M.C.M. acknowledges support from the Simons Foun-
dation. M.C.M. and X.Y. acknowledge support from NSF-
DMR-305184. The authors also acknowledge the Syracuse
University HTC Campus Grid, NSF award ACI-1341006 and
the Soft Matter Program at Syracuse University.
■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●����������������������������������������■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●����������������������������������������■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■■●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●●����������������������������������������Dr=0.01Dr=1Dr=100[4] Julio M. Belmonte, Gilberto L. Thomas, Leonardo G. Brun-
net, Rita M. C. de Almeida, and Hugues Chat´e. Self-propelled
particle model for cell-sorting phenomena. Phys. Rev. Lett.,
100:248702, Jun 2008.
[5] E. Ben-Isaac, ´E. Fodor, P. Visco, F. van Wijland, and Nir S.
Gov. Modeling the dynamics of a tracer particle in an elastic
active gel. Phys. Rev. E, 92:012716, Jul 2015.
[6] Ludovic Berthier. Nonequilibrium glassy dynamics of self-
propelled hard disks. Physical Review Letters, 112(22):220602,
2014.
[7] Dapeng Bi, J. H. Lopez, J. M. Schwarz, and M. Lisa Manning.
A density-independent rigidity transition in biological tissues.
Nat Phys, advance online publication:–, 09 2015.
[8] Dapeng Bi, Jorge H. Lopez, J. M. Schwarz, and M. Lisa Man-
ning. Energy barriers and cell migration in densely packed tis-
sues. Soft Matter, 10:1885–1890, 2014.
[9] Fabio Cecconi, Andrea Puglisi, Umberto Marini Bettolo Mar-
coni, and Angelo Vulpiani. Noise activated granular dynamics.
Phys. Rev. Lett., 90:064301, Feb 2003.
[1] A. R. Abate and D. J. Durian. Effective temperatures and
activated dynamics for a two-dimensional air-driven granu-
lar system on two approaches to jamming. Phys. Rev. Lett.,
101:245701, Dec 2008.
[2] Thomas E. Angelini, Edouard Hannezo, Xavier Trepat, Jef-
frey J. Fredberg, and David A. Weitz. Cell migration driven
by cooperative substrate deformation patterns. Phys. Rev. Lett.,
104:168104, Apr 2010.
[3] Thomas E. Angelini, Edouard Hannezo, Xavier Trepat, Manuel
Marquez, Jeffrey J. Fredberg, and David A. Weitz. Glass-like
dynamics of collective cell migration. Proceedings of the Na-
tional Academy of Sciences, 108(12):4714–4719, 2011.
[10] M. Deforet, V. Hakim, H. G. Yevick, G. Duclos, and P. Sil-
berzan. Emergence of collective modes and tri-dimensional
structures from epithelial confinement. Nat Commun, 5, 05
2014.
[11] Raphal Etournay, Marko Popovi, Matthias Merkel, Amitabha
Nandi, Corinna Blasse, Benot Aigouy, Holger Brandl, Gene
Myers, Guillaume Salbreux, Frank Jlicher, and Suzanne Eaton.
Interplay of cell dynamics and epithelial tension during mor-
phogenesis of the Drosophila pupal wing. eLife, 4:e07090, jun
2015.
[12] Reza Farhadifar,
Jens-Christian Roeper, Benoit Aigouy,
Suzanne Eaton, and Frank Julicher. The influence of cell me-
chanics, cell-cell interactions, and proliferation on epithelial
packing. Current Biology, 17(24):2095 – 2104, 2007.
[13] Yaouen Fily, Silke Henkes, and M. Cristina Marchetti. Freez-
ing and phase separation of self-propelled disks. Soft Matter,
10:2132–2140, 2014.
[14] Yaouen Fily and M. Cristina Marchetti. Athermal phase sepa-
ration of self-propelled particles with no alignment. Phys. Rev.
Lett., 108:235702, Jun 2012.
[15] Simon Garcia, Edouard Hannezo, Jens Elgeti, Jean-Franois
Joanny, Pascal Silberzan, and Nir S. Gov. Physics of active
jamming during collective cellular motion in a monolayer. Pro-
ceedings of the National Academy of Sciences, 112(50):15314–
15319, 2015.
[16] Michael A. Gelbart, Bing He, Adam C. Martin, Stephan Y.
Thiberge, Eric F. Wieschaus, and Matthias Kaschube. Volume
conservation principle involved in cell lengthening and nucleus
movement during tissue morphogenesis. Proceedings of the Na-
tional Academy of Sciences, 109(47):19298–19303, 2012.
11
[17] N.P.A.Devika Gunasinghe, Alan Wells, ErikW. Thompson, and
HonorJ. Hugo. Mesenchymalepithelial transition (met) as a
mechanism for metastatic colonisation in breast cancer. Can-
cer and Metastasis Reviews, 31(3-4):469–478, 2012.
[18] Anna Haeger, Marina Krause, Katarina Wolf, and Peter Friedl.
Cell jamming: Collective invasion of mesenchymal tumor cells
imposed by tissue confinement. Biochimica et Biophysica Acta
(BBA) - General Subjects, 1840(8):2386 – 2395, 2014. Matrix-
mediated cell behaviour and properties.
[19] S. Henkes, M. van Hecke, and W. van Saarloos. Critical jam-
ming of frictional grains in the generalized isostaticity picture.
EPL (Europhysics Letters), 90(1):14003, 2010.
[20] Silke Henkes, Yaouen Fily, and M. Cristina Marchetti. Active
jamming: Self-propelled soft particles at high density. Phys.
Rev. E, 84:040301, Oct 2011.
[21] Hisao Honda. Description of cellular patterns by dirichlet do-
mains: The two-dimensional case. Journal of Theoretical Biol-
ogy, 72(3):523 – 543, 1978.
[22] Lars Hufnagel, Aurelio A. Teleman, H. Rouault, Stephen M.
Cohen, and Boris I. Shraiman. On the mechanism of wing size
determination in fly development. Proceedings of the National
Academy of Sciences, 104(10):3835–3840, March 2007.
[23] Atsushi Ikeda, Ludovic Berthier, and Peter Sollich. Unified
study of glass and jamming rheology in soft particle systems.
Phys. Rev. Lett., 109:018301, Jul 2012.
[24] Alexandre J. Kabla. Collective cell migration: leadership, in-
vasion and segregation. Journal of The Royal Society Interface,
9(77):3268–3278, 2012.
[25] Karen E Kasza, Dene L Farrell, and Jennifer A Zallen.
Spatiotemporal control of epithelial
remodeling by regu-
lated myosin phosphorylation. Proc Natl Acad Sci U S A,
111(32):11732–11737, Aug 2014.
[26] Jae Hun Kim, Xavier Serra-Picamal, Dhananjay T. Tambe,
Enhua H. Zhou, Chan Young Park, Monirosadat Sadati, Jin-
Ah Park, Ramaswamy Krishnan, Bomi Gweon, Emil Millet,
James P. Butler, Xavier Trepat, and Jeffrey J. Fredberg. Propul-
sion and navigation within the advancing monolayer sheet. Nat
Mater, 12(9):856–863, 09 2013.
[27] G. Lejeune Dirichlet. ber die reduction der positiven quadratis-
chen formen mit drei unbestimmten ganzen zahlen. Journal fur
die reine und angewandte Mathematik, 40:209–227, 1850.
[28] Bo Li and Sean X. Sun. Coherent motions in confluent cell
monolayer sheets. Biophysical Journal, 107(7):1532–1541,
2015/08/23 2015.
[29] Andrea J. Liu and Sidney R. Nagel. The jamming transition
and the marginally jammed solid. Annual Review of Condensed
Matter Physics, 1(1):347–369, 2010.
[30] M. L. Manning and A. J. Liu. Vibrational modes identify
soft spots in a sheared disordered packing. Phys. Rev. Lett.,
107:108302, Aug 2011.
[31] M. Lisa Manning, Ramsey A. Foty, Malcolm S. Steinberg, and
Eva-Maria Schoetz. Coaction of intercellular adhesion and cor-
tical tension specifies tissue surface tension. Proceedings of
the National Academy of Sciences, 107(28):12517–12522, 07
2010.
[32] Cecile Monthus and Jean-Philippe Bouchaud. Models of traps
and glass phenomenology. Journal of Physics A: Mathematical
and General, 29(14):3847, 1996.
[33] Tatsuzo Nagai and Hisao Honda. A dynamic cell model for the
formation of epithelial tissues. Philosophical Magazine Part B,
81(7):699–719, 2001.
[34] Yukiko Nakaya, Shinya Kuroda, Yuji T. Katagiri, Kozo
Kaibuchi, and Yoshiko Takahashi. Mesenchymal-epithelial
transition during somitic segmentation is regulated by differen-
tial roles of cdc42 and rac1. Developmental Cell, 7(3):425–438,
2015.
[35] Ran Ni, Martien A. Cohen Stuart, and Marjolein Dijkstra. Push-
ing the glass transition towards random close packing using
self-propelled hard spheres. Nat Commun, 4, 10 2013.
[36] Kenechukwu David Nnetu, Melanie Knorr, Josef Kas, and
Mareike Zink. The impact of jamming on boundaries of collec-
tively moving weak-interacting cells. New Journal of Physics,
14(11):115012, 2012.
[37] Peter Olsson and S. Teitel. Athermal jamming versus thermal-
ized glassiness in sheared frictionless particles. Phys. Rev. E,
88:010301, Jul 2013.
[38] Jin-Ah Park, Jae Hun Kim, Dapeng Bi, Jennifer A. Mitchel,
Nader Taheri Qazvini, Kelan Tantisira, Chan Young Park, Mau-
reen McGill, Sae-Hoon Kim, Bomi Gweon, Jacob Notbohm,
Robert Steward Jr, Stephanie Burger, Scott H. Randell, Alvin T.
Kho, Dhananjay T. Tambe, Corey Hardin, Stephanie A. Shore,
Elliot Israel, David A. Weitz, Daniel J. Tschumperlin, Eliza-
beth P. Henske, Scott T. Weiss, M. Lisa Manning, James P. But-
ler, Jeffrey M. Drazen, and Jeffrey J. Fredberg. Unjamming
and cell shape in the asthmatic airway epithelium. Nat Mater,
14(10):1040–1048, 10 2015.
[39] L Petitjean, M Reffay, E Grasland-Mongrain, M Poujade,
B Ladoux, A Buguin, and P Silberzan. Velocity fields in
journal,
a collectively migrating epithelium.
98(9):1790–1800, 2010.
Biophysical
[40] M. Poujade, E. Grasland-Mongrain, A. Hertzog, J. Jouanneau,
P. Chavrier, B. Ladoux, A. Buguin, and P. Silberzan. Col-
lective migration of an epithelial monolayer in response to a
model wound. Proceedings of the National Academy of Sci-
ences, 104(41):15988–15993, 10 2007.
[41] Alberto Puliafito, Lars Hufnagel, Pierre Neveu, Sebastian Stre-
ichan, Alex Sigal, D. Kuchnir Fygenson, and Boris I. Shraiman.
Collective and single cell behavior in epithelial contact inhi-
bition. Proceedings of the National Academy of Sciences,
109(3):739–744, 2012.
[42] Monirosadat Sadati, Nader Taheri Qazvini, Ramaswamy Kr-
ishnan, Chan Young Park, and Jeffrey J. Fredberg. Collective
migration and cell jamming. Differentiation, 86(3):121 – 125,
2013. Mechanotransduction.
[43] E.-M. Schoetz, R. D. Burdine, F. Julicher, M. S. Steinberg, C.-
P. Heisenberg, and R. A. Foty. Quantitative differences in tis-
sue surface tension influence zebrafish germlayer positioning.
HFSP journal, Vol.2 (1):1–56, 2008.
[44] Eva-Maria Schoetz, Marcos Lanio, Jared A. Talbot, and M. Lisa
Manning. Glassy dynamics in three-dimensional embryonic tis-
sues. J. Roy. Soc. Interface, 10:20130726, 2013.
[45] Nestor Sepulveda, Laurence Petitjean, Olivier Cochet, Erwan
Grasland-Mongrain, Pascal Silberzan, and Vincent Hakim. Col-
lective cell motion in an epithelial sheet can be quantitatively
described by a stochastic interacting particle model. PLoS Com-
put Biol, 9(3):e1002944, 03 2013.
[46] Leonardo E. Silbert. Jamming of frictional spheres and random
loose packing. Soft Matter, 6:2918–2924, 2010.
12
[47] Peter Sollich. Rheological constitutive equation for a model of
soft glassy materials. Phys. Rev. E, 58:738–759, Jul 1998.
[48] S S Soumya, Animesh Gupta, Andrea Cugno, Luca Deseri,
Kaushik Dayal, Dibyendu Das, Shamik Sen, and Mandar M.
Inamdar. Coherent motion of monolayer sheets under confine-
ment and its pathological implications. PLoS Comput Biol,
11(12):e1004670, 12 2015.
[49] D. B Staple, R Farhadifar, J. C Roeper, B Aigouy, S Eaton,
and F Julicher. Mechanics and remodelling of cell packings in
epithelia. Eur. Phys. J. E, 33(2):117–127, Oct 17 2010.
[50] A Szabo, Runnep, E Mehes, W O Twal, W S Argraves, Y Cao,
and A Czirok. Collective cell motion in endothelial monolayers.
Physical Biology, 7(4):046007, 2010.
[51] B. Szabo, G. J. Szollosi, B. Gonci, Zs. Jur´anyi, D. Selmeczi,
and Tam´as Vicsek. Phase transition in the collective migration
of tissue cells: Experiment and model. Phys. Rev. E, 74:061908,
Dec 2006.
[52] Jean Paul Thiery. Epithelial-mesenchymal transitions in tumour
progression. Nat Rev Cancer, 2(6):442–454, 06 2002.
[53] Erik W. Thompson and Donald F. Newgreen. Carcinoma inva-
sion and metastasis: A role for epithelial-mesenchymal transi-
tion? Cancer Research, 65(14):5991–5995, 2005.
[54] S Torquato, Author and HW Haslach, Jr. Random heteroge-
neous materials: Microstructure and macroscopic properties.
Applied Mechanics Reviews, 55(4):B62–B63, 07 2002.
[55] V Trappe, V Prasad, L Cipelletti, and P Segre. . . . Jamming
phase diagram for attractive particles. Nature, Jan 2001.
[56] L´eon Van Hove. Correlations in space and time and born ap-
proximation scattering in systems of interacting particles. Phys.
Rev., 95:249–262, Jul 1954.
[57] Tam´as Vicsek, Andr´as Czir´ok, Eshel Ben-Jacob, Inon Cohen,
and Ofer Shochet. Novel type of phase transition in a system
of self-driven particles. Phys. Rev. Lett., 75:1226–1229, Aug
1995.
[58] Denis L Weaire and Stefan Hutzler. The physics of foams. Ox-
ford University Press, 1999.
[59] F. Wetzel, A. Fritsch, D. Bi, R. Stange, S. Pawlizak, T. Kieβling,
L-C. Horn, K. Bendrat, M. Oktay, M. Zink, A. Niendorf, J. Con-
deelis, M. Hockel, M. C. Marchetti, M. L. Manning, and J. K.
Kas. Unpublished, 2015.
[60] Ian Y. Wong, Sarah Javaid, Elisabeth A. Wong, Sinem Perk,
Daniel A. Haber, Mehmet Toner, and Daniel Irimia. Collective
and individual migration following the epithelial–mesenchymal
transition. Nat Mater, 13(11):1063–1071, 11 2014.
[61] N. Xu, V. Vitelli, A. J. Liu, and S. R. Nagel. Anharmonic and
quasi-localized vibrations in jammed solidsmodes for mechan-
ical failure. EPL (Europhysics Letters), 90(5):56001, 2010.
[62] X. B. Yang, D. Bi, M. C. Marchetti, and M. L. Manning. Un-
published, 2015.
[63] Steven M. Zehnder, Melanie Suaris, Madisonclaire M. Bellaire,
and Thomas E. Angelini. Cell volume fluctuations in {MDCK}
monolayers. Biophysical Journal, 108(2):247 – 250, 2015.
[64] Steven M. Zehnder, Marina K. Wiatt, Juan M. Uruena, Al-
ison C. Dunn, W. Gregory Sawyer, and Thomas E. An-
gelini. Multicellular density fluctuations in epithelial monolay-
ers. Phys. Rev. E, 92:032729, Sep 2015.
|
1507.02774 | 1 | 1507 | 2015-07-10T03:48:46 | Multivalent ion-mediated nucleic acid helix-helix interactions: RNA versus DNA | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech",
"physics.comp-ph"
] | Ion-mediated interaction is critical to the structure and stability of nucleic acids. Recent experiments suggest that the multivalent ion-induced aggregation of double-stranded (ds) RNAs and DNAs may strongly depend on the topological nature of helices, while there is still lack of an understanding on the relevant ion-mediated interactions at atomistic level. In this work, we have directly calculated the potentials of mean force (PMF) between two dsRNAs and between two dsDNAs in Cobalt Hexammine ion (Co-Hex) solutions by the atomistic molecular dynamics simulations. Our calculations show that at low [Co-Hex], the PMFs between B-DNAs and between A-RNAs are both (strongly) repulsive.However, at high [Co-Hex], the PMF between B-DNAs is strongly attractive, while those between A-RNAs and between A-DNAs are still (weakly) repulsive. The microscopic analyses show that for A-form helices, Co-Hex would become internal binding into the deep major groove and consequently cannot form the evident ion-bridge between adjacent helices, while for B-form helices without deep grooves, Co-Hex would exhibit external binding to strongly bridge adjacent helices. In addition, our further calculations show that, the PMF between A-RNAs could become strongly attractive either at very high [Co-Hex] or when the bottom of deep major groove is fixed with a layer of water. | physics.bio-ph | physics | Multivalent ion-mediated nucleic acid helix-helix interactions:
RNA versus DNA
Yuan-Yan Wu1, Zhong-Liang Zhang1, Jin-Si Zhang1, Xiao-Long Zhu2 and Zhi-Jie Tan1*
1Department of Physics and Key Laboratory of Artificial Micro & Nano-structures of Ministry of
Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China
2Department of Physics, School of Physics & Information Engineering, Jianghan University, Wuhan 430056,
China
Abstract
Ion-mediated interaction is critical to the structure and stability of nucleic acids. Recent
experiments suggest that the multivalent ion-induced aggregation of double-stranded (ds) RNAs
and DNAs may strongly depend on the topological nature of helices, while there is still lack of an
understanding on the relevant ion-mediated interactions at atomistic level. In this work, we have
directly calculated the potentials of mean force (PMF) between two dsRNAs and between two
3+ (Co-Hex) solutions by the atomistic molecular dynamics simulations. Our
dsDNAs in Co(NH3)6
calculations show that at low [Co-Hex], the PMFs between B-DNAs and between A-RNAs are
both (strongly) repulsive. However, at high [Co-Hex], the PMF between B-DNAs is strongly
attractive, while those between A-RNAs and between A-DNAs are still (weakly) repulsive. The
microscopic analyses show that for A-form helices, Co-Hex would become “internal binding” into
the deep major groove and consequently cannot form the evident ion-bridge between adjacent
helices, while for B-form helices without deep grooves, Co-Hex would exhibit “external binding”
to strongly bridge adjacent helices. In addition, our further calculations show that, the PMF
between A-RNAs could become strongly attractive either at very high [Co-Hex] or when the
bottom of deep major groove is fixed with a layer of water.
Key words: RNA, DNA, potential of mean force, ion-binding, atomistic molecular dynamics
*To whom correspondence should be addressed: [email protected]
1
1. Introduction
Nucleic acids are highly charged polyanions (1). The structural folding of nucleic acids into
compact native structure generally experiences strong Coulomb repulsive force, while metal ions
in solution would bind to nucleic acid surface to reduce the Coulomb repulsion during folding
(2-11). Therefore, metal ions play a critical role in the folding thermodynamics and kinetics of
nucleic acids (2-11).
The double-stranded (ds) helix is a fundamental segment in nucleic acid structures (1). The
ion-mediated interaction between two ds helices would provide the energetics for RNA structural
collapse and DNA condensation (1-11). For dsDNAs, the existing experiments suggest that
monovalent ions can only mediate an inter-helix repulsion, while multivalent ions can induce an
attractive force and consequently cause DNA condensation (2,4-6,12-19). Such multivalent
ion-mediated effective attraction has also been proposed to cause the condensation of other
like-charged polyelectrolytes (20-25).
Due to the similarity between dsDNA and dsRNA in charge density (1), it is natural to assume
that the ion-mediated force between dsRNAs would be similar to that between dsDNAs, and short
dsDNAs have been used instead of dsRNAs to probe the ion-dependent structural assembly of
nucleic acid helices (26-29) despite their different helix structures. However, the recent UV
adsorption experiments have shown that in 20 mM Na+ buffer, ~4mM Cobalt hexamine
3+, i.e., Co-Hex) ions can cause the aggregation of short dsDNAs, but cannot lead to the
(Co(NH3)6
aggregation of short dsRNAs (30,31). The remarkable finding is beyond the expectation and has
been proposed to be attributed to the binding of Co-Hex ions into deep major groove of RNAs
(30-32). However, there is still lack of a direct illustration on the linkage between ion-mediated
interaction and the relevant ion-binding structure, corresponding to the experimental ionic
conditions (30-32). In this work, we will directly calculate the potentials of mean force between
two dsRNAs as well as those between two dsDNAs, in order to directly establish the relationship
between the ion-mediated effective force and the ion-binding structures.
Several polyelectrolyte theories have been developed and employed to predict the
ion-mediated interaction between like-charged polyelectrolytes. The counterion condensation
theory has been developed based on line-charge structural model of infinite length, while the
theory always predicts effective attractive forces between dsDNAs even at low monovalent salt and
is only applicable to line-charge polyelectrolytes (33,34). The Poisson-Boltzmann theory with
mean-field assumption ignores the ion-ion correlations and thus always predicts effective repulsive
2
force between dsDNAs even at high multivalent salt (35-41). The electrostatic zipper model can
predict an effective force between two helices at multivalent salt while the partition of binding ions
into major/minor grooves is somewhat ad hoc (42,43). The tightly bound ion model which
accounts for ion-ion correlation can predict the ion-mediated effective force between dsDNAs,
while the model assumes the distribution of molecular charges on phosphate groups and thus may
not make reliable predictions on ion-binding structures near helix grooves (44-47). Therefore, the
existing theories could not give an ab initio and direct illustration on the multivalent ion-mediated
effective interaction between dsRNAs (and dsDNAs) and the microscopic ion-binding structure. As
an important bridge between experiments and theories, the computer simulations can be a powerful
tool to probe the effective interactions between biomolecules (48-57). Beyond the simplified
Monte Carlo and Langevin dynamics simulations (e.g., 47-50), the all-atomistic molecular
dynamics (MD) explicitly takes into account the detailed atomistic structure of nucleic acids, ions,
and water molecules, and thus could give a direct and detailed exploration on the effective
interactions between biomolecules (48-63).
In this study, we will employ the all-atomistic MD simulations to directly calculate the
potentials of mean force (PMF) between A-form RNAs (A-RNAs) and those between B-form
DNAs (B-DNAs) in Co-Hex salt solutions. We will also make further calculations for A-form
DNAs (A-DNAs) and for the A-RNAs with the spatial ion-accessible region modified.
Correspondingly, the detailed Co-Hex binding structures will be analyzed. We will also make the
direct comparisons with the related experiments. The present work would give atomistic level
calculations on the effective interactions between dsRNAs as well as those between dsDNAs, and
would present a direct illustration on the relationship between ion-mediated effective interactions
and ion-binding structures.
2. Model and method
2.1 All-atomistic molecular dynamics simulations
In the work, we will calculate the PMFs between two A-RNAs, between two B-DNAs and
between two A-DNAs by the atomistic MD simulations. The A-RNA, B-DNA and A-DNA helices
are of 16-bp length, and their atomic structures are displayed in Fig. 1. The sequences of the
helices are selected according to the recent experiments (30-32) and contain all the dinucleotide
base pairs; see Table S1 in Supplementary Material.
The structures of the 16-bp nucleic acid helices are taken as the standard A-RNA, B-DNA and
A-DNA fibers. The two parallel A-RNA (or DNA) helices with axes in z direction are separated in
3
x direction and immersed in a rectangle box containing explicit water and ions. The A-RNAs (or
DNAs) are harmonically restrained with 1000 kJ/(mol.nm2) force constant in y and z directions,
thus are only allowed to move translationally in x direction. The size of the rectangle box is taken
3+ (Co-Hex) ions are added with Amber tleap
as 130Å×80Å×100 Å. The Na+ and Co(NH3)6
Program (64,65). To get the desirable bulk ion concentrations, the simplified Monte Carlo
simulations (50,66) are employed to estimate numbers of Co-Hex and Na+ ions in the simulational
cell before the all-atomistic MD simulations; see Supplementary Material for details. Afterwards,
the numbers of Co-Hex and Na+ ions from the simplified Monte Carlo simulations are used in the
all-atomistic simulations, and the realistic bulk Na+ and Co-Hex concentrations from the
all-atomistic MD simulations are close to the desirable values; see Fig. S1 in Supplementary
Material for the cases of 100mM Na+ and 5mM Co-Hex solutions.
In the simulations, we employ the Amber parmbsc0 force field and the TIP3P water model
combined with parmbsc0 ion model (64,65). Corresponding to the recent experiments (30-32), the
solutions always contain ~100mM NaCl as the background, and [Co-Hex] is taken as low (0.5mM),
high (5mM) and very high (50mM) values, respectively. Since previous experiments showed that
the octahedral coordination shell remains intact in binding to DNA (67), the Co-Hex ions are built
with the explicit bonds between cobalt and amine groups and the explicit N–Co–N angles specified
to generate an octahedral Co-Hex complex (68,69). The charges on a Co-Hex complex are
generated based on the electrostatic potential generated with the DGAUSS program (68), and the
hydrogen bonds of Co-Hex are considered with van der Waals potential with the new charge model
and improved van der Waals parameters (68,70). All the parameters for Co-Hex ions are taken
from Ref. (68). All the systems are optimized, thermalized (298K) and equilibrated by the program
Gromacs 4.5 (71) with the periodic boundary conditions and Particle Mesh Ewald method
employed (72), and a time step of 1-2fs is used in the conjunction with LINCS algorithm (73).
Firstly, an energy minimization of 5,000 steps is performed with the steepest descent algorithm at
low temperature, and then the systems are slowly heated to 298K and equilibrated with the
Nose-Hoover temperature coupling until 0.5 ns (74). Afterwards, the subsequent NPT simulations
of 2 ns (time-step 1fs, P=1atm) are performed with the Parrinello-Rahman pressure coupling and
with the nucleic acids fixed (65). Finally, each MD simulation is continued for another 60ns in the
isothermic-isobaric ensemble (time-step 2fs, P=1atm, T=298K). Our MD simulations generally
reach the equilibrium after ~10ns, as shown in Figs. S2 and S3 for ion-binding number and Figs.
S4 and S5 for helix-helix separation versus MD time in Supplementary Material. The trajectories in
equilibrium are used to calculate the PMF between two helices.
4
2.2 Calculating potential of mean force between two helices
In the work, we employ the pseudo-spring method (53,54,66) to calculate the PMF between
two A-RNAs (or DNAs). In the method, a pseudo-spring with spring constant k is added to link the
center of mass of two helices in MD simulations, as shown in Fig. 1. Based on the MD trajectories
in equilibrium, the effective force between the two A-RNA (or DNA) helices can be calculated by
, (1)
where Δx is the deviation of the spring length away from the original length x0 in equilibrium. The
negative and positive Δx’s correspond to the attractive and repulsive forces, respectively. After a
series of F(x) at different separations x are obtained, the PMF ΔG(x) between the two A-RNA (or
DNA) helices can be calculated through the integration (53,54,66)
, (2)
where xref is the outer reference separation. It has been shown previously that the pseudo-spring
method is efficient and convenient in calculating PMF between two DNAs and two like-charged
nanoparticles (53,54,66). In practice, the spring constant k is generally taken as 1000 kJ/(mol.nm2)
and xref is taken as 40Å. For the cases that two helices interact very strongly, we also use a higher
value of k=2000 kJ/(mol·nm2). Additionally, we have also made the additional calculations for the
PMFs with the umbrella sampling method (70,71,75), and the results are very close to those from
the pseudo-spring method; see Fig. S6 in Supplementary Material.
3. Results and discussion
In the work, we will calculate the PMF ΔG(x) between two nucleic acid helices in Co-Hex
solutions by the all-atomistic MD simulations, and will examine how Co-Hex ions modulate the
PMFs between dsDNAs and dsRNAs. We would emphasize illustrating the microscopic
mechanism for the difference in PMFs between A-RNAs and B-DNAs.
3.1 PMFs between B-DNAs at low and high [Co-Hex]s
As shown in Fig. 2a, the PMF between two B-DNAs is repulsive at low (~0.5mM) [Co-Hex],
while becomes strongly attractive with the free energy minimum of ~-3.5kBT at the axis-axis
separation of ~27Å when [Co-Hex] is increased to ~5mM. Such ion-mediated PMF is generally
coupled to ion-binding (45,76). Due to the high entropy penalty for Co-Hex binding at low
[Co-Hex], the system at low [Co-Hex] is dominated by Na+-binding from the background of
100mM Na+ and the monovalent ions can only modulate repulsive PMF. In contrast, at high
5
Fkxrefref()()()(')'xxGxGxGxFxdx[Co-Hex], Co-Hex-binding would dominate the system due to lower entropy ion-binding penalty,
and strong Coulomb attraction between Co-Hex and two adjacent B-DNAs could cause an
attractive force. The predicted PMFs between B-DNAs are in accordance with the previous
experiments which show that high [Co-Hex] could induce DNA aggregation while DNAs resist
condensation at low [Co-Hex] (12-17,30,31). The axis-axis separation of ~27Å at the lowest ΔG(x)
for ~5mM [Co-Hex] also agrees well with the values of equilibrium spacing of DNA aggregates
from various experiments (15-19,77). In addition, following Refs. (46,78), we have calculated the
osmotic pressures for hexagonal DNA aggregate, with assuming the additivity for the pair-wise
PMF (28). The calculated osmotic pressures are very close to the corresponding experimental data
(15,16); see Fig. S8 and Supplementary Material for details.
To gain a deep understanding on the relationship between [Co-Hex] and the PMF between
B-DNAs, we would analyze the radial Co-Hex concentration distributions c(r)’s around B-DNAs
at different [Co-Hex]’s, and c(r)’s have been calculated according to Eq. 9 in Ref (79). As shown in
Fig. 2b, Co-Hex ions would begin to bind to a B-DNA at axial distance>~5Å, and prefer to
accumulate at the outer surface of B-DNA with the axial distance of ~13Å. Such binding near the
outer surface of a helix with the radial distance range of [11-15Å] will be termed as “external
binding” (31). Higher (~5mM) [Co-Hex] causes much more Co-Hex ions “external binding”
around B-DNAs near the radial distance of ~13Å than low (~0.5mM) [Co-Hex]. The further
detailed analyses show that at low [Co-Hex], Co-Hex ions prefer to bind over the minor groove
rather than the major groove, while at high [Co-Hex], Co-Hex ions mainly bind over the major
groove rather than the minor groove; see Fig. 2c. This is reasonable since the minor groove of
B-DNA is narrower and electrically more negative, thus Co-Hex ions prefer to binding to minor
groove at low [Co-Hex] (31). When [Co-Hex] is increased to a high value, more Co-Hex ions
become binding while the narrow minor groove has already accommodated many binding ions,
causing the binding of excess Co-Hex ions over the major groove. The strong “external binding” of
Co-Hex would be shared by adjacent B-DNAs, and could cause an significantly attractive PMF
between B-DNAs at high [Co-Hex] (46,53,55,66,80).
To analyze the 3-dimensional Co-Hex binding around two B-DNAs, we would illustrate the
binding structure of Co-Hex ions of high concentration (74). As shown in Figs. 3ab, Co-Hex ions
form the obvious ion-bridge configuration between adjacent B-DNAs which appears much more
pronounced for high [Co-Hex]. Such apparent ion-bridge configuration between two helices would
provide a key contribution to the attractive PMF between B-DNAs (46,50,53,55,66).
At atomistic level, we have analyzed the structure of water molecules around bridging Co-Hex
6
between two DNAs when two DNAs strongly attract each other. As shown in Fig. S9 in
Supplementary Material, bridging Co-Hex can induce the ordering of water molecules between
adjacent phosphates in two DNAs, i.e., Co-Hex can induce the rotation of H2O with O atoms
pointing to Co-Hex and H atoms pointing to phosphates. Such configuration of bridging
Co-Hex-induced ordering of water molecules between two helices could mediate an apparent
DNA-DNA attraction, as suggested by Parsegian, Rau, Qiu and coworkers (13,19).
3.2 PMFs between A-RNAs at low, high and very high [Co-Hex]s
The PMFs between A-RNAs have been calculated at low (~0.5mM), high (~5mM) and very
high (~50mM) [Co-Hex]’s, as shown in Fig. 2d. At low (~0.5mM) [Co-Hex], in analogy to
B-DNAs, the PMF between A-RNAs is repulsive and appears slightly stronger than that between
B-DNAs. This may be attributed to the slightly higher charge density and thicker helix of A-RNAs
(1). As [Co-Hex] is increased to ~5mM, the PMF between A-RNAs is still (weakly) repulsive with
weaker strength than that at low [Co-Hex], which is distinctively in contrast to the strongly
attractive PMF between B-DNAs. With the increase of [Co-Hex] to a very high value (~50mM),
the PMF between A-RNAs becomes strongly attractive with free energy minimum of ~-4.1kBT at
axis-axis separation of ~27Å. The predicted PMFs at different [Co-Hex]’s are in good accordance
with the recent experiments on dsRNAs which have shown that dsRNAs resist condensation when
[Co-Hex] <~5mM while could condense at very high [Co-Hex], relatively to the background Na+
(30,31).
To understand the [Co-Hex]-dependent PMF between A-RNAs, we have analyzed the radial
Co-Hex concentration distribution around A-RNAs (79), as shown in Figs. 2ef. For convenience,
corresponding to the above termed “external binding” with the radial distance range of [11-15Å]
(31), another binding mode with the radial distance range of <~11Å (~helical radii of dsDNA and
dsRNA) is termed as “internal binding” (31). Overall, Co-Hex ions begin to bind to A-RNA at
radial distance >~2 Å, which is attributed to the accessible deep major groove of A-RNA. This is
distinctly different from B-DNA. At low (~0.5mM) [Co-Hex], Co-Hex ions exhibit “internal
binding” into the deep major groove within the radial distance of <~10Å. As [Co-Hex] is increased
to ~5mM, Co-Hex ions prefer to bind internally into the deep major groove around the radial
distance of ~8Å, and Co-Hex binding distribution is extended to the radial distance range of
<~12Å. With the increase of [Co-Hex] to ~50mM, Co-Hex ions would still show preference to
bind internally into the major groove around radial distance of 8-9Å, while the apparent binding
distribution is extended to the radial distance range of <~16Å and the [Co-Hex] at ~15Å can nearly
reach ~0.3M. The concentration-dependent Co-Hex-binding distribution around A-RNA is
7
understandable. At low [Co-Hex], Co-Hex would prefer to bind into the deep major groove with
small radial distance where electric field is strongest (31), and as [Co-Hex] is increased, more
Co-Hex ions become binding and begin to externally bind near outer surface of A-RNA after the
deep/narrow major groove is fulfilled. Very high (e.g., ~50mM) [Co-Hex] would cause apparent
“external binding”, and such apparent “external binding” of Co-Hex can be shared by adjacent
A-RNAs to cause an attractive PMF at very high [Co-Hex].
The 3-dimensional Co-Hex binding around two A-RNAs has been directly illustrated in Figs.
3e-f. At low [Co-Hex], Co-Hex ions of high density are “internal binding” in the deep and narrow
major groove, and there is no visible ion-bridge between two A-RNA surfaces. At high (~5mM)
[Co-Hex], the abundant Co-Hex ions of high density reside in the major groove, and there is still
no evident ion-bridge, which corresponds to the (weakly) repulsive PMF between A-RNAs. When
[Co-Hex] is increased to ~50mM, many more Co-Hex ions become binding and the excess Co-Hex
would bind externally at outer surface of A-RNAs, and the apparent ion-bridge between two
A-RNAs is formed. Such ion-bridging configuration with Co-Hex-induced water ordering would
be responsible for the strongly attractive PMF between A-RNAs at very high [Co-Hex]
(46,50,53,55,66).
3.3. PMF is dependent on helical structure: dsRNA versus dsDNA
3.3.1. A-RNA versus B-DNA
Both A-RNAs and B-DNAs are highly (negatively) charged polymers, while the effective
interaction between A-RNAs appears distinctly different from that between B-DNAs at high
(~5mM) [Co-Hex], i.e., the PMF between B-DNAs is strongly attractive while that between
A-RNAs is repulsive; see Fig. 4a and Figs. 2ad. As shown above, there are distinct differences
between A-RNA and B-DNA in their topological structures. A-RNA has the deep/narrow major
groove and the wide/shallow minor groove, as shown in Fig. 3. Since there is strongest electric
potential in the deep/narrow major groove of A-RNA, Co-Hex would prefer to bind deeply into
major groove. Such “internal binding” cannot contribute to the formation of ion-bridge between
two A-RNAs and consequently the PMF is repulsive. Unlike A-RNA structure, B-DNA has the
wide major groove and the narrow minor groove which are not deep compared with those of
A-RNA, and Co-Hex ions would like to become “external binding” near outer surface of B-DNA
rather than “internal binding” deeply into grooves. Such “external binding” would help to form the
apparent ion-bridge which could be shared by adjacent B-DNAs to result in an attractive PMF.
When [Co-Hex] is increased to a very high value, the excess Co-Hex can also “externally bind” to
A-RNA since the deep/narrow major groove is fulfilled by more binding Co-Hex ions. Such
8
“external binding” induced by a very high [Co-Hex] can also promote the formation of ion-bridge
and cause a strong attractive force between A-RNAs.
3.3.2. PMF between A-DNAs
To further confirm the above described effects of “internal binding” and “external binding”, we
have calculated and analyzed the PMF between two A-DNAs at 5mM [Co-Hex]. As shown in Fig.
4a, in analogy to A-RNA, the PMF between A-DNAs is repulsive at 5mM [Co-Hex]. The detailed
analysis on Co-Hex binding distribution also shows the “internal binding” near the radial distance
of ~8Å, which is similar to that of A-RNA; see Figs. 4bc. As illustrated in Fig. 3c, such “internal
binding” is in deep major groove and resists the formation of ion-bridge, which is responsible for
the (weakly) repulsive PMF between A-DNAs. It is noted here that despite the similarity in
“internal binding”, the radial distribution of Co-Hex around A-DNA is slightly different from that
around A-RNA: Co-Hex concentration around A-RNA is higher at radial distance of <~5 Å, while
lower at radial distance of ~8 Å than that around A-DNA; see Figs. 4bc. Such difference comes
from the difference in helical structures of A-RNA and A-DNA. Despite the similar A-form helical
structure, as indicated in the experiments (81-83), the width of major groove of A-DNA is ~2Å
narrower than that of A-RNA, which is also in accordance with our used structure parameters. The
narrower major groove of A-DNA would cause Co-Hex ions to have more preference to bind in
major groove and coordinate with adjacent phosphate strands at the radial distance of ~8Å, which
has been observed in our MD simulations and illustrated in Fig. 3c. The radial profiles of Co-Hex
concentrations around A-form helices are in accordance with the previous experiments on Co-Hex
binding to A-DNA, which showed that Co-Hex would bind to bases in major groove and to
phosphates, either bridging across narrow major groove or residing between two adjacent
intra-strand phosphates (84).
3.3.3. PMF between “modified” A-RNAs
Since it is the “internal binding” that resists the formation of ion-bridge and causes the
repulsive PMF between A-RNAs, there would be an interesting question: If Co-Hex ions are
prohibited to enter deeply into the major groove of A-RNAs, can A-RNAs attract each other? To
answer the interesting question and further validate the above analysis, we have made the further
calculations for the “modified” A-RNAs (A-RNA’ and A-RNA”) at 5mM [Co-Hex]. A-RNA’
corresponds to the A-RNA with the bottom of the central 1/3 major groove fixed by a layer of
water, and A-RNA” corresponds to the A-RNA with the bottom of the whole major groove fixed by
a layer of water, as shown in Fig. 5. Such treatment of fixing water at the bottom of deep major
groove may effectively exclude the very deep binding of Co-Hex in major groove, and may
9
possibly cause the “external binding” and strong attractive force between the modified A-RNAs.
Our calculations show that there are indeed apparently attractive forces between the “modified”
A-RNAs; see Fig. 6a. The analyses on Co-Hex concentration distributions around the “modified”
A-RNAs also show that the treatment of excluding Co-Hex from the deep major groove promotes
the “external binding”. As shown in Figs. 6bc, when a layer of water is fixed at the bottom of major
groove, the original “internal binding” in the radial distance range of ~2.5Å to ~12.5 Å would
change into the binding in the radial distance range of ~7Å to ~12.5 Å, and the “external binding”
near the radial distance of ~13Å. Furthermore, the direct illustration in Fig. 5 shows that such
treatment would definitely help to form the ion-bridge between two A-RNAs, which could cause
the strong attractive PMF between the modified A-RNAs (46,50,53,55,66).
4. Conclusion
In this work, the atomistic MD simulations have been employed to calculate and analyze the
potentials of mean force between A-RNAs, between B-DNAs, and between A-DNAs in Co-Hex
solutions. The present work shows that though the nucleic acid helices have similar negative
charge densities, the effective interactions between them are distinctively different. At high
[Co-Hex], two B-DNAs strongly attract each other, while two A-RNAs and two A-DNAs repel
each other. The present analysis shows that such significant difference between B-form and A-form
helices is attributed to the ion-binding structure. For B-DNA, Co-Hex ions would become
“external binding” around phosphates and form the ion-bridge between two B-DNAs. But for
A-RNA and A-DNA, due to the existence of deep major groove, Co-Hex would preferentially
exhibit “internal binding” into the major groove and consequently cannot form the ion-bridge
between two A-form helices, causing the repulsive interaction between them. The effective
interactions between A-RNAs can become strongly attractive when Co-Hex ions become “external
binding” to form the apparent ion-bridge, which can be realized by either increasing Co-Hex to a
very high concentration or fixing a layer of water at the bottom of the deep major groove.
Overall, our results are in accordance with the experimental findings (17,30,31,77). First, our
calculations show that the PMF between A-RNAs is weakly repulsive while that between B-DNAs
is strongly attractive at ~5mM Co-Hex, which is in good agreement with the recent experiments
(30,31). Second, Parsegian, Rau and coworkers combined single-molecule magnetic tweezers with
osmotic stress on DNA assembly in various salt solutions. Their direct measurement of the free
energy for DNA aggregates in Co-Hex solution is ~-0.21 kBT/bp (17), a value close to ours of
~-3.5kBT for 16-bp DNAs, and the separation of ~27Å at the lowest free energy for ~5mM Co-Hex
from our MD simulations is also close to the experimental equilibrium spacing ~28 Å of DNA
10
aggregate (15,17). Third, the calculated osmotic pressures with assuming additivity of PMF are
very close to the corresponding experimental measurements (15,16).
Kornyshev and Leikin have successfully developed their electrostatic zipper model to explain
why A-RNAs resist condensation while B-DNAs would become condensed, and such distinctive
behaviors were proposed to be attributed to the different widths of the major grooves of different
helices (42). However, the model has also involved some important simplifications such as
assuming the uniform distribution of ions in grooves and ignoring ion size and 3-dimensional
ion-accessible geometry of different helices. The present all-atom MD simulations with explicit
helical structures, ions and water molecules, show that the 3-dimensional topology of different
helices can play an important role in modulating PMF between two helices. Specifically, the
attractive PMF between B-DNA at ~5mM Co-Hex is accompanied with the external binding of
Co-Hex above major groove at radial distance of ~13Å, while the non-attractive PMF between
A-RNAs at ~5mM Co-Hex is accompanied with the internal binding of Co-Hex at radial distance
of ~8-9Å in major groove.
The present work has also involved some assumptions and simplifications. First, all the
dsRNAs and dsDNAs have been approximately treated as rigid bodies and thus the stability and
flexibility of ds helices have been ignored. Since all the nucleic acid helices are in solutions of
100mM NaCl, the helices would have high stability to keep its helical structure rather than become
denatured (85,86). Although nucleic acid helices are flexible in ionic solutions, the ignorance of
helix flexibility can be a reasonable approximation due to the high persistence length (45nm-60nm)
(81,82) and strong stretching modulus (>~500pN) (74,87,88) of dsDNA and dsRNA. Second, we
have made the calculations for A-DNAs without considering the likely dependence of A-DNA
structure on ionic condition (1). It is a reasonable simplified treatment since we only use A-DNA as
a structure model to further examine the mechanism. Thirdly, the extensive relevant experiments
have used the 25-bp dsRNA and dsDNA helices (30,31), while 16-bp dsRNAs and dsDNAs have
been employed in the work for saving the computational time. Such simplification would not affect
the conclusions since the experiments have shown that 16-bp dsDNAs have the similar
Co-Hex-dependent condensation behavior (30). Fourthly, the related experiments involve different
background Na+ concentrations of 100mM and 20mM with different experiment techniques
(30,31), while the present work only considers the solutions containing 100mM Na+. Such
simplification should not qualitatively affect our conclusions since the experiments with different
background Na+ concentrations have shown the qualitatively similar results (30,31). The effect of
competition between Co-Hex and background Na+ on the effective nucleic acid helix-helix
interaction may deserve to be studied separately.
11
Furthermore, in the present work, we only consider the parallel configuration of two helices
with all the atoms restrained in y and z directions, and the helices cannot rotate around any of the x,
y and z axes. In fact, such parallel configuration would be the most favorable one for large
separation between two helices (89), while may become unfavorable for the closely packaging of
two helices, due to the strong electrostatic repulsion (28,29). In realistic 3D space, the axes of two
helices at small separation can rotate by a small angle to form a typical X-shaped structure (90,91)
with several possible packaging modes for different helical structures (90,91). Such tight
non-parallel helix-helix configuration may be important for the packaging of some RNAs such as
the P4-P6 domain of the Tetrahymena thermophila intron (92).
Finally, the present work only involves the system of two helices while the related experiments
generally involved multiple helices. The previous studies have shown that the multi-body effect
may slightly affect the multivalent-mediated helix-helix attraction and consequently may slightly
affect the comparisons between our predictions and experiments on the interaction strength and
axis-axis separation (28). The strict and extensive exploration for multi-body PMF between helices
at atomistic level is deserved to be studied separately.
Nevertheless, the present work has provided the direct calculations on the PMFs between
A-RNAs, between B-DNAs and between A-DNAs, and has directly illustrated the microscopic
mechanisms for the different PMFs between double-stranded nucleic acids with different helical
structures.
5. Acknowledgements
We are grateful to Profs. Shi-Jie Chen, Haiping Fang and Wenbing Zhang for valuable
discussions, and Chang Shu for facility assistance. This work was supported by the National Key
Scientific Program (973)-Nanoscience and Nanotechnology (No. 2011CB933600), the National
Science Foundation of China grants (11074191, 11175132 and 11374234), and the Program for
New Century Excellent Talents (Grant No. NCET 08-0408). One of us (Y.Y. Wu) also thanks the
financial support from the interdisciplinary and postgraduate programs under the Fundamental
Research Funds for the Central Universities.
12
References
1. Bloomfield, V.A., Crothers, D.M. and Tinoco, I.Jr. (2000) Nucleic Acids: Structures, Properties, and
Functions. University science books.
2. Bloomfield, V.A. (1997) DNA Condensation by multivalent cations. Biopolymers 44, 269-282.
3. Chen, S.J. (2008) RNA folding: conformational statistics, folding kinetics, and ion electrostatics. Annu. Rev.
Biophys. 37, 197.
4. Woodson, S.A. (2010) Compact intermediates in RNA folding. Annu. Rev. Biophys. 39, 61-77.
5. Wong, G.C., and Pollack, L. (2010) Electrostatics of strongly charged biological polymers: ion-mediated
interactions and self-organization in nucleic acids and proteins. Annu. Rev. Phys. Chem. 61, 171-89.
6. Auffinger, P., Grover, N., Westhof, E. (2011) Metal ion binding to RNA[J]. Met. Ions Life Sci., 9, 1-35.
7. Draper, D.E. (2013) Folding of RNA tertiary structure: linkages between backbone phosphates, ions, and
water. Biopolymers. 99, 1105-13.
8. Lipfert, J., Doniach, S., Das, R., and Herschlag, D. (2014) Understanding nucleic acid-ion interactions. Annu.
Rev. Biochem., 83, 813-41.
9. Koculi, E., Hyeon, C., and Thirumalai, D., Woodson, S.A. (2007) Charge density of divalent metal cations
determines RNA stability. J. Am. Chem. Soc., 129, 2676-82.
10. Tan, Z.J., and Chen, S.J. (2011) Importance of diffuse metal ion binding to RNA. Met. Ions Life Sci., 9,
101-124.
11. Williams, L.D., Maher III, L.J. (2000) Electrostatic mechanisms of DNA deformation. Annu. Rev. Biophys.
Biomol. Struct., 29, 497-521.
12. Widom, J., and Baldwin, R.L. (1980) Cation-induced toroidal condensation of DNA studies with Co3+(NH3)6.
J. Mol. Biol., 144, 431-53.
13. DeRouchey, J., Parsegian, V.A., and Rau, D.C. (2010) Cation charge dependence of the forces driving DNA
assembly. Biophys. J., 99, 2608-15.
14. Korolev, N., Berezhnoy, N.V., Eom, K.D., Tam, J.P., and Nordenskiöld, L.(2012) A universal description for
the experimental behavior of salt-(in)dependent oligocation-induced DNA condensation. Nucleic Acids Res.,
40, 2808-21.
15. Rau, D.C., and Parsegian, V.A. (1992) Direct measurement of temperature-dependent solvation forces
between DNA double helices. Biophys. J., 61, 260-71.
16. Rau, D.C., and Parsegian,V.A.(1992) Direct measurement of
the
intermolecular forces between
counterion-condensed DNA double helices. Evidence for long range attractive hydration forces. Biophys. J.,
61, 246-59.
17. Todd, B.A., Parsegian, V.A., Shirahata, A., Thomas, T.J., and Rau, D.C. (2008) Attractive forces between
cation condensed DNA double helices. Biophys. J., 94, 4775-82.
18. Qiu, X., Andresen, K., Kwok, L.W., Lamb, J.S., Park, H.Y., and Pollack, L. (2007) Inter-DNA Attraction
Mediated by Divalent Counterions. Phys. Rev. Lett., 99, 038104.
19. Qiu, X., Parsegian, V. A., and Rau, D.C. (2010) Divalent Counterion-induced condensation of triple-strand
DNA. Proc. Natl. Acad. Sci. USA, 107, 21482-21486.
20. Butler, J. C., Angelini, T., Tang, J.X., and Wong, G.C. (2003) Ion multivalence and like-charge polyelectrolyte
attraction. Phys. Rev. Lett., 91, 028301.
13
21. Wu, J.Z., Bratko, D., and Prausnitz, J.M. (1998) Interaction between like-charged colloidal spheres in
electrolyte solutions. Proc. Natl. Acad. Sci. USA, 95, 15169-15172.
22. Solis, F.J. and Olvera de la Cruz, M. (2000) Collapse of flexible polyelectrolytes in multivalent salt solutions.
J. Chem. Phys., 112, 2030-2035.
23. Angelini, T. E., Liang, H., Wriggers, W., and Wong, G.C. (2003) Like-charge attraction between
polyelectrolytes induced by counterion charge density waves. Proc. Natl. Acad. Sci. USA, 100, 8634-8637.
24. Tang, J. X., Janmey, P.A., Lyubartsev, A., and Nordenskiöld, L. (2002) Metal ion-induced lateral aggregation
of filamentous viruses fd and m13. Biophys. J., 83, 566-581.
25. Wen, Q., and Tang, J.X. (2006) Temperature effects on threshold counterion concentration to induce
aggregation of fd virus. Phys. Rev. Lett., 97, 048101.
26. Bai, Y., Das, R., Millett, I.S., Herschlag, D. and Doniach, S. (2005) Probing counterion modulated repulsion
and attraction between nucleic acid duplexes in solution. Proc. Natl Acad. Sci. USA, 102, 1035-1040.
27. Bai, Y., Chu, V.B., Lipfert, J., Pande, V.S., Herschlag, D., and Doniach S. (2008) Critical assessment of
nucleic acid electrostatics via experimental and computational investigation of an unfolded state ensemble. J.
Am. Chem. Soc., 130, 12334-41.
28. Tan, Z.J., and Chen, S.J. (2006) Electrostatic free energy landscapes for nucleic acid helix assembly. Nucleic
Acids Res., 34, 6629-6639.
29. Tan, Z.J., and Chen, S.J. (2012) Ion-mediated RNA structural collapse: effect of spatial confinement. Biophys.
J., 103, 827-836.
30. Li, L., Pabit, S.A., Meisburger, S.P., and Pollack, L. (2011) Double-stranded RNA resists condensation, Phys.
Rev. Lett., 106, 108101.
31. Tolokh, I.S., Pabit, S.A., Katz, A.M., Chen, Y., Drozdetski, A., Baker, N., Pollack. L., and Onufriev, A.V.
(2014) Why double-stranded RNA resists condensation. Nucleic Acids Res., 42, 10823-31.
32. Pabit, S.A., Qiu, X., Lamb, J.S., Li, L., Meisburger, S.P., and Pollack, L. (2009) Both helix topology and
counterion distribution contribute to the more effective charge screening in dsRNA compared with dsDNA.
Nucleic Acids Res., 37, 3887-96.
33. Manning, G. S. (1978) The Molecular theory of polyelectrolyte solutions with applications to the electrostatic
properties of polynucleotides. Q. Rev. Biophys., 11, 179-246.
34. Ray, J., and Manning, G. S.. (1994) An attractive force between two rodlike polyions mediated by the sharing
of condensed counterions. Langmuir., 10, 2450–2461.
35. Baker, N.A., Sept, D., Joseph, S., Holst, M.J., and McCammon, J.A. (2001) Electrostatics of nanosystems:
Application to microtubules and the ribosome. Proc. Natl. Acad. Sci. USA, 98, 10037-10041.
36. Boschitsch, A. H., Fenley, M. O., and Zhou, H.. (2002) Fast boundary element method for the linear
Poisson-Boltzmann equation. J. Phys. Chem. B, 106, 2741-2754.
37. Baker, N. A. (2004) Poisson–Boltzmann methods for biomolecular electrostatics. Method. Enzymol., 383,
94-118.
38. Lu, B., Cheng, X., Huang, J., and McCammon, J.A. (2010) AFMPB: an adaptive fast multipole
Poisson–Boltzmann solver for calculating electrostatics in biomolecular systems. Comput. Phys. Commun.,
181, 1150-1160.
39. Chen, D., Chen, Z., Chen, C., Geng, W., and Wei, G.W. (2011) MIBPB: a software package for electrostatic
analysis. J. Comput. Chem., 32, 756-70.
40. Wei, G.W., Zheng, Q., Chen, Z., and Xia, K. (2012) Variational multiscale models for charge transport. SIAM
Rev., 54, 699-754.
14
41. Ovanesyan, Z., Medasani, B., Fenley, M.O., Guerrero-García, G.I., Olvera de la Cruz, M., and Marucho, M.
(2014) Excluded volume and ion-ion correlation effects on the ionic atmosphere around B-DNA: Theory,
simulations, and experiments. J. Chem. Phys., 141, 225103.
42. Kornyshev, A.A., and Leikin, S. (2013) Helical structure determines different susceptibilities of dsDNA,
dsRNA, and tsDNA to counterion-induced condensation. Biophys. J., 104:2031-41.
43. Kornyshev, A.A. and Leikin, S. (1999) Electrostatic zipper motif for DNA aggregation. Phys. Rev. Lett., 82,
4138.
44. Tan, Z.J., and Chen, S.J. (2005) Electrostatic correlations and fluctuations for ion binding to a finite length
polyelectrolyte. J. Chem. Phys., 122, 044903.
45. Tan, Z.J., and Chen, S.J. (2010) Predicting ion binding properties for RNA tertiary structures. Biophys. J., 99,
1565-1576.
46. Tan, Z.J., and Chen, S.J. (2006) Ion-mediated nucleic acid helix-helix Interactions. Biophys. J., 91, 518-536.
47. Tan, Z.J., and Chen, S.J. (2011) Salt contribution to RNA tertiary structure folding stability. Biophys. J., 101,
176-187.
48. Lyubartsev, A.P., and Nordenskiold, L. (1995) Monte Carlo simulation study of ion distribution and osmotic
pressure in hexagonally oriented DNA. J. Phys. Chem., 99, 10373–10382.
49. E. Allahyarov, Gompper, G., and Löwen, H.. (2004) Attraction between DNA molecules mediated by
multivalent ions. Phys. Rev. E, 69, 041904.
50. Wang, F.H., Wu, Y.Y., and Tan, Z.J. (2013) Salt contribution to the flexibility of single-stranded nucleic acid
of finite length. Biopolymers, 99, 370-81.
51. Shi, Y.Z., Wang, F.H., Wu, Y.Y., and Tan Z.J. (2014) A coarse-grained model with implicit salt for RNAs:
predicting 3D structure, stability and salt effect. J. Chem. Phys., 141, 105102.
52. Feig, M. and Pettitt, B.M. (1998) Structural equilibrium of DNA represented with different force fields.
Biophys. J., 75, 134-149.
53. Luan, B., and Aksimentiev, A. (2008) DNA attraction in monovalent and divalent electrolytes. J Am Chem
Soc., 130, 15754-5.
54. Maffeo, C., Luan, B., and Aksimentiev, A. (2012) End-to-end attraction of duplex DNA. Nucleic Acids Res.,
40, 3812-21.
55. Dai, L., Mu, Y., Nordenskiöld, L., and van der Maarel, J.R. (2008) Molecular dynamics simulation of
multivalent-ion mediated attraction between DNA molecules. Phys. Rev. Lett., 100, 118301.
56. Anandakrishnan R, Drozdetski A, Walker RC, Onufriev AV. (2015) Speed of conformational change:
comparing explicit and implicit solvent molecular dynamics simulations. Biophys. J., 108, 1153-64.
57. Anandakrishnan, R., Aguilar, B. and Onufriev, A. V. (2012). H++ 3.0: automating pK prediction and the
preparation of biomolecular structures for atomistic molecular modeling and simulations. Nucleic Acids Res.,
40(W1), W537-W541.
58. Yu, T., Wang, X. Q., Sang, J. P., Pan, C. X., Zou, X. W., Chen, T. Y. and Zou, X. (2012). Influences of
mutations on the electrostatic binding free energies of chloride ions in Escherichia coli ClC. J. Phys. Chem. B,
116, 6431-6438.
59. Wang, X. Q., Yu, T., Sang, J. P., Zou, X. W., Chen, T. Y., Bolser, D. and Zou, X. (2010). A three-state
multi-ion kinetic model for conduction properties of ClC-0 chloride channel. Biophys. J., 99, 464-471.
60. Zhang, Y., Zhang, J., and Wang, W. (2011) Atomistic analysis of pseudoknotted RNA unfolding. J. Am. Chem.
Soc., 133, 6882-6885.
15
61. Li, W., Nordenskiöld, L., Zhou, R., and Mu, Y. (2014) Conformation-dependent DNA attraction. Nanoscale, 6,
7085-92.
62. Gong, Z., Zhao, Y., Chen, C, Duan, Y. and Xiao, Y. (2014) Insights into ligand binding to PreQ1 riboswitch
aptamer from molecular dynamics simulations. PloS one, 9, e92247.
63. Auffinger, P., Westhof, E. (1998) Simulations of the molecular dynamics of nucleic acids. Curr. Opin. Struc.
Boil., 8, 227-236.
64. Pérez, A., Marchán, I., Svozil, D., Sponer, J., Cheatham, T. E. III, Laughton, C.A., and Orozco, M. (2007)
Refinement of the AMBER Force Field for Nucleic Acids: Improving the Description of< i> α/< i> γ
Conformers. Biophys. J., 92, 3817-3829.
65. Van Der Spoel, D., Lindahl, E., Hess, B., Groenhof, G., Mark, A.E. and Berendsen, H.J. (2005) GROMACS:
fast, flexible, and free. J. Comput. Chem., 26, 1701-1718.
66. Wu, Y.Y., Wang, F.H., and Tan, Z.J. (2013) Calculating potential of mean force between like-charged
nanoparticles: a comprehensive study on salt effects. Phys. Lett. A, 377, 1911-1919.
67. Braunlin W. H., Anderson C. F., Record Jr M. T. (1987) Competitive interactions of cobalt (3+) hexamine and
sodium with helical B-DNA probed by cobalt-59 and sodium-23 NMR. Biochemistry, 26, 7724-7731.
68. Cheatham T. E., Kollman P. A. (1997) Insight into the stabilization of A-DNA by specific ion association:
spontaneous B-DNA
to A-DNA
transitions observed
in molecular dynamics simulations of d
[ACCCGCGGGT]2 in the presence of hexaamminecobalt (III). Structure, 5, 1297-1311.
69. Marques H. M, Brown K. L. (2002) Molecular mechanics and molecular dynamics simulations of porphyrins,
metalloporphyrins, heme proteins and cobalt corrinoids. Coord. Chem. Rev., 225, 123-158.
70. Galindo-Murillo R., Cheatham T. E. (2014) DNA Binding Dynamics and Energetics of Cobalt, Nickel, and
Copper Metallopeptides. Chem. Med. Chem., 9, 1252-1259.
71. Hess, B., Kutzner, C., Van Der Spoel, D., and Lindahl, E. (2008) GROMACS 4: Algorithms for highly
efficient, load-balanced, and scalable molecular simulation. J. Chem. Theory Comput., 4, 435-447.
72. Essmann, U., Perera, L., Berkowitz, M. L., Darden, T., Lee, H. and Pedersen, L. G. (1995) A smooth particle
mesh Ewald method. J. Chem. Phys., 103, 8577-8593.
73. Hess, B., Bekker, H., Berendsen, H. J. and Fraaije, J. G. (1997) LINCS: a linear constraint solver for
molecular simulations. J. Comput. Chem., 18(12), 1463-1472.
74. Wu, Y.Y., Bao, L., Zhang, X., and Tan, Z.J. (2015) Flexibility of short DNAs with finite-length effect: from
base pairs to tens of base pairs. J. Chem. Phys., 142:125103(1-13), 2015.
75. Kästner J and Thiel, W. (2005) Bridging the gap between thermodynamic integration and umbrella sampling
provides a novel analysis method: “Umbrella integration”. J. Chem. Phys., 123, 144104.
76. Qiu, X., Giannini, J., Howell, S.C., Xia, Q., Ke, F., and Andresen, K. (2013) Ion competition in condensed
DNA arrays in the attractive regime. Biophys. J., 105:984-92.
77. Qiu, X., Rau, D.C., Parsegian, V.A., Fang, L.T., Knobler, C.M., and Gelbart, W.M. (2011) Salt-dependent
DNA-DNA spacings in intact bacteriophage λ reflect relative importance of DNA self-repulsion and bending
energies. Phys. Rev. Lett., 106, 028102.
78. He, Z., Chen, S. J. (2013) Quantifying Coulombic and solvent polarization-mediated forces between DNA
helices. J. Phys. Chem. B 117, 7221-7
79. Kirmizialtin, S., Silalahi, A.R., Elber, R., and Fenley, M.O. (2012) The ionic atmosphere around A-RNA:
Poisson-Boltzmann and molecular dynamics simulations. Biophys. J., 102:829-38.
80. Alsharif, S.A., Chen, L.Y., Tlahuice-Flores, A., Whetten, R.L. and Yacaman, M.J. (2014) Interaction between
functionalized gold nanoparticles in physiological saline. Phys. Chem. Chem. Phys., 16, 3909-3913.
16
81. McKee, T., and McKee, J. R. (2009) Biochemistry: the molecular basis of life. Oxford University Press.
82. Weeks, K.M., and Crothers, D. M. (1993) Major groove accessibility of RNA. Science, 261, 1574-1574.
83. Shakked, Z., Rabinovich, D., Cruse, W.B.T, Egert, E., Kennard, O., Sala, G, S. Salisbury, A., and Viswamitra,
M.A.. (1981) Crystalline A-DNA: the X-ray analysis of the fragment d (GGTATACC). Proc. Royal. Soc.
London B: Biol. Sci., 213, 479-487.
84. Gao, Y.G., Robinson, H., van Boom, J.H., Wang, A.H. (1995) Influence of counter-ions on the crystal
structures of DNA decamers: binding of [Co(NH3)6]3+ and Ba2+ to A-DNA. Biophys. J., 69, 559-68.
85. Tan, Z.J., and Chen, S.J. (2007) RNA helix stability in mixed Na+/Mg2+ solution. Biophys. J., 92, 3615-32.
86. Tan, Z.J., and Chen, S.J. (2006) Nucleic acid helix stability: effects of salt concentration, cation valence and
size, and chain length. Biophys. J., 90, 1175-90.
87. Lipfert, J., Skinner, G.M., Keegstra, J.M., Hensgens, T., Jager, T., Dulin, D., Köber, M., Yu, Z., Donkers, S.P.,
Chou, F.C., Das, R., and Dekker, N.H. (2014) Double-stranded RNA under force and torque: similarities to
and striking differences from double-stranded DNA. Proc. Natl. Acad. Sci. USA, 111, 15408-13.
88. Herrero-Galán, E., Fuentes-Perez, M.E., Carrasco, C., Valpuestam J.M., Carrascosam J.L., Moreno-Herrero,
F., and Arias-Gonzalez, J.R. (2013) Mechanical identities of RNA and DNA double helices unveiled at the
single-molecule level. J. Am. Chem. Soc., 135, 122-31.
89. Kornyshev, A. A., Leikin, S. (2000) Electrostatic interaction between long, rigid helical macromolecules aat ll
interaxial angles. Phys. Rev. E, 62, 2576-96.
90. Murthy, V. L., Rose, G. D. (2000) Is counterion delocalization responsible for collapse in RNA folding?
Biochemistry. 39, 14365-70.
91. Várnai, P., Timsit, Y. (2010) Differential stability of DNA crossovers in solution mediated by divalent cations.
Nucleic Acids Res., 38, 4163-72.
92. Cate, J. H., Gooding, A. R, Podell, E., Zhou, K., Golden, B. L., Kundrot, C. E., Cech, T. R., Doudna, J. A.
(1996) Crystal structure of a group I ribozyme domain: principles of RNA packing. Science, 273, 1678-85.
17
Figure 1 An illustration for two parallel 16-bp A-RNAs (a), B-DNAs (b) and A-DNAs (c). The
spring with a spring constant k which connects the centers of mass of two helices has been used to
calculate the potential of mean force between two double-stranded RNAs and DNAs (53,54,73).
18
Figure 2 (a) The potentials of mean force between two 16-bp B-DNAs in 0.5mM and 5mM
Co-Hex solutions. (b) The Co-Hex ion distributions around the B-DNA corresponding to panel a.
(c) The Co-Hex ion distributions around B-DNA in (or over) major groove and minor groove
according to panels a and b. (d) The potentials of mean force between two 16-bp A-RNAs in
0.5mM, 5mM and 50mM Co-Hex solutions. (e) The Co-Hex distributions around the A-RNA
corresponding to panel d. (c) The Co-Hex distributions around A-RNA in (or over) major groove
and minor groove corresponding to panels d and e. Note that the buffers always contain 100mM
NaCl.
19
Figure 3 (a,b) The illustrations for the region of high Co-Hex ion charge density (larger than 0.02
e/Å3) around two 16-bp B-DNAs in 0.5mM (a) and 5mM (b) Co-Hex solutions; (c) The illustration
for the region of the high Co-Hex ion charge density (larger than 0.02 e/Å3) around two 16-bp
A-DNAs in 5mM Co-Hex solutions. (d-f) The illustrations for the region of high Co-Hex charge
density (larger than 0.02 e/Å3) around two 16-bp A-RNAs in 0.5mM (d), 5mM (e) and 50mM (f)
Co-Hex solutions. Note that the buffers always contain 100mM NaCl.
20
Figure 4 (a) The potentials of mean force between two 16-bp nucleic acid (B-DNA, A-RNA, and
A-DNA) helices in 5mM Co-Hex ion solution; (b) The Co-Hex distribution around the nucleic acid
helices corresponding to panel a; (c) The Co-Hex distribution around the nucleic acid helices in (or
over) major groove and minor groove corresponding to panels a and b. Note that the buffers always
contain 100mM NaCl.
21
Figure 5 The illustrations for the spatially modified A-RNA (A-RNA’ and A-RNA”) and the region
of the high Co-Hex charge density (larger than 0.02 e/Å3) in 5mM Co-Hex solution around two
16-bp A-RNA’s and around two 16-bp A-RNA’s. The A-RNA’ denotes the A-RNA with the bottom
of the central 1/3 major groove fixed with a layer of water, whereas the A-RNA” denotes the
A-RNA with the bottom of the entire major groove fixed with a layer of water. The red-gray chains
in deep major grooves illustrate the fixed water molecules. Note that the buffers always contain
100mM NaCl.
22
Figure 6 (a) The potentials of mean force between two 16-bp A-RNAs, between two 16-bp
A-RNA’s, and between 16-bp A-RNA’’s in 5mM Co-Hex ion solution; (b) The Co-Hex
distributions around the A-RNA, A-RNA’, and A-RNA’’ corresponding to panels a and b; (c) The
Co-Hex distributions around the A-RNA, A-RNA’, and A-RNA’’ in (or over) major groove and
minor groove corresponding to panels a and b. The A-RNA’ denotes the A-RNA with the bottom of
the central 1/3 major groove fixed with a layer of water, whereas the A-RNA” denotes the A-RNA
with the bottom of the entire major groove fixed with a layer of water. Note that the buffers always
contain 100mM NaCl.
23
Supplementary Material for
Multivalent ion-mediated nucleic acid helix-helix interactions:
RNA versus DNA
Yuan-Yan Wu1, Zhong-Liang Zhang1, Jin-Si Zhang1, Xiao-Long Zhu2 and Zhi-Jie Tan1*
1Department of Physics and Key Laboratory of Artificial Micro & Nano-structures of Ministry of
Education, School of Physics and Technology, Wuhan University, Wuhan 430072, China
2Department of Physics, School of Physics & Information Engineering, Jianghan University, Wuhan 430056,
China
*To whom correspondence should be addressed: [email protected]
24
To get desirable bulk Co-Hex and Na+ concentrations
Due to the competition between monovalent and multivalent ions in binding to nucleic acids
(e.g., Ref. (29)), it is not straightforward to obtain desirable bulk monovalent/multivalent ion
concentrations in a MD simulation for a nucleic acid in a mixed monovalent/multivalent ion solution.
In the present work, to get the desirable bulk ion concentrations, before the all-atom MD
simulations, the simplified Monte Carlo (MC) simulations (50,66) are employed to estimate numbers
of Co-Hex and Na+ ions in the simulational cell. Practically, all-atom structure of nucleic acids and
ions are placed in the MC simulational cell, and water is treated as continuous medium with
dielectric constant of 78 (50,66). The MC algorithm with Coulomb and Lenard-Jones potentials
(50,66) is performed to get the bulk ion concentrations in equilibrium. We change the relative
numbers of Co-Hex and Na+ ions and repeat the MC processes, and then we can estimate the
numbers of Co-Hex and Na+ ions in the simulational cell at the desirable bulk ion concentrations.
Afterwards, the numbers of Co-Hex and Na+ ions from the simplified MC simulations are used in the
all-atom simulations, and the realistic bulk Na+ and Co-Hex concentrations from the all-atom MD
simulations are very close to the desirable values; see Fig. S1 in Supplementary Material for the
cases of 100mM Na+/5mM Co-Hex solutions.
25
Osmotic pressure for DNA aggregates and comparison with experimental data
For DNA array (DNA aggregates), the osmotic pressures have been measured experimentally by
the osmotic stress technique. In this section, based on the pair-wise DNA helix-helix interactions, we
have calculated the osmotic pressures and compared the results with the experimental data. For a
hexagonal DNA array (see Fig. 11a in Ref. (46)), the mean free energy Δg(x) per DNA can be
approximately calculated through the summation over the pair-wise helix-helix interactions between
nearest neighbor pairs (46,78)
, (S1)
where
is the total free energy between an helix and its six neighbors, and the
factor 1/2 is used to remove double-counting. ΔGi(x) is the free energy for the two-helix system of a
helix and its i-th neighbor. Here, we have neglected the nonadditive effect in multiple helix packing
and kept only interactions between the nearest neighboring helices, i.e., when calculating the
interaction between two helices, we ignore the existence of other helices.
The osmotic pressure П(x) as a function of the helix-helix distance x can be calculated from
(46,78)
, (S2)
where V=L×A is the volume of the hexagonal region around each helix. L is the length of each helix,
and
A=3 is the average cross section area per molecule in the DNA array (46). Practically,
we first fit the calculated PMF to a polynomial function and based on the fitted PMF, we could easily
calculate the osmotic pressures according to Eqs. S1 and S2.
As shown in Fig. S8, for DNAs in 5mM Co-Hex/100mM Na+, the calculated osmotic pressures
are very close to the experimental data (15,16), and the slight deviation may come from the
multi-helix effect (46) which was ignored in our calculations for hexagonal DNA aggregate. For
DNAs in 0.5mM Co-Hex/100mM Na+, there is no directly available experimental data. As shown in
Fig. S8, the addition of 0.5mM Co-Hex in 100mM Na+ can cause a different osmotic pressure curve
from that for 100mM Na+ (Ref. (1) in supplementary material).
26
612/)()(iixGxg61)()(ixGxgVxgx)()(2/32xTable S1 A-RNA, B-DNA and A-DNA sequences used in the present study.a
Nucleic Acids
Sequences
A-RNA
B-DNA
A-DNA
5’-CGACUCUACUACGCGC-3’
GCUGAGAUGAUGCGCG
5’-CGACTCTACTACGCGC-3’
GCTGAGATGATGCGCG
5’-CGACTCTACTACGCGC-3’
GCTGAGATGATGCGCG
aThe sequences of the short RNA and DNAs are selected according to the recent small angle
scattering experiments and UV measurements (29).
27
Figure S1 (a,c) The realistic bulk Co-Hex (a) and Na+ (c) concentrations for two DNAs in 100mM
Na+/5mM Co-Hex solutions from the all-atom MD simulations as a function of DNA-DNA distance;
(b,d) The realistic bulk Co-Hex (b) and Na+ (d) concentrations for two RNAs in 100mM Na+/5mM
Co-Hex solutions from the all-atom MD simulations as a function of RNA-RNA distance.
28
Figure S2 Number of bound Co-Hex ions within the distances of 11Å (red) and 16Å (green) from
two duplex helical axes as function of MD running time: (a) B-DNAs and (b) A-RNAs at 5mM
Co-Hex and 100mM Na+ ion solution. Dashed lines denote the averaged values in equilibrium.
29
Figure S3 Number of bound Na+ ions within the distances of 11Å (red) and 16Å (green) from two
duplex helical axes as function of MD running time: (a) B-DNA and (b) A-RNA at 5mM Co-Hex and
100mM Na+ ion solution. Dashed lines denote the averaged values in equilibrium.
30
Figure S4 An illustration of the MD convergence of separation x between the centers of mass of two
16-bp A-RNA helices in 5mM Co-Hex ion solution with 100mM NaCl. The averaged values of
separation x are made over every Δt by
and Δt =20ps. The black lines denote the
averaged values in equilibrium. Two typical original separations x0 between two A-RNAs are shown
in the panels.
31
tdttxtttt/')'(2/2/
Figure S5 An illustration of the MD convergence of separation x between the centers of mass of two
16-bp B-DNA helices in 5mM Co-Hex ion solution with 100mM NaCl. The averaged values of
separation x are made over every Δt by
and Δt =20ps. The black lines denote the
averaged values in equilibrium. Two typical original separations x0 between two A-DNAs are shown
in the panels.
32
tdttxtttt/')'(2/2/
Figure S6 The potential of mean force as a function of the separation x between the centers of mass
of two 16-bp nucleic acid helices in 5mM Co-Hex ion solution with 100mM NaCl: (a) A-RNAs and
(b) B-DNAs. Blue: calculated from the umbrella sampling with the weighted histogram analysis
method (WHAM); Red: calculated from the pseudo-spring method employed in the present work.
33
Figure S7 The Co-Hex charge distribution per unit charge on nucleic acid helices as a function of
distance r from the axes of nucleic acid helices. (a) B-DNAs at 0.5mM and 5mM [Co-Hex]’s; (b)
A-RNAs at 0.5mM, 5mM and 50 mM [Co-Hex]’s; (c) The comparisons between A-RNAs, B-DNA
sand A-DNAs; (d) The modified A-RNAs. The A-RNA’ denotes the A-RNA with the bottom of the
central 1/3 major groove fixed with a layer of water, whereas the A-RNA” denotes the A-RNA with
the bottom of the entire major groove fixed with a layer of water. Please note that the buffers always
contain 100mM NaCl.
34
Figure S8 Comparison of osmotic pressure curves for hexagonal DNA aggregate in various salt
solutions from MD simulations (lines) and experiments (symbols). Lines: DNAs at 0.5mM (green)
and 5mM (blue) [Co-Hex]’s, and the buffers always contain 100mM NaCl. Symbols: ▼,
experimental data for DNAs at 5mM Co-Hex and 100mM NaCl with 10mM TrisCl (17); ●,
experimental data for DNAs at 100mM NaCl with 10mM TrisCl (Ref. 1 in Supplementary Material).
The apparent deviation between blue curve and ● comes from that the experiments did not involve
Co-Hex ions.
35
Figure S9 Two snapshots to illustrate the structure of water molecules which link the bridging
Co-Hex and phosphates in two DNAs. The axis-axis distances between DNAs are 30Å (a) and 28 Å
(b), respectively. The dash lines are H-bonds which are automatically displayed by the software
VMD (Ref. 2 in Supplementary Material).
36
References
93. Rau, D. C, Lee, B., Parsegian, V. A. (1984) Measurement of the repulsive force between polyelectrolyte
molecules in ionic solution: hydration forces between parallel DNA double helices. Proc. Natl. Acad. Sci.
USA 81, 2621-5.
94. Humphrey, W., Dalke, A., Schulten K. (1996) VMD: visual molecular dynamics. J. Mol. Graphics, 14, 33-38.
37
|
1607.00120 | 1 | 1607 | 2016-07-01T06:28:23 | Holographic intravital microscopy for 2-D and 3-D imaging intact circulating blood cells in microcapillaries of live mice | [
"physics.bio-ph",
"physics.med-ph",
"physics.optics",
"q-bio.QM"
] | Intravital microscopy is an essential tool that reveals behaviours of live cells under conditions close to natural physiological states. So far, although various approaches for imaging cells in vivo have been proposed, most require the use of labelling and also provide only qualitative imaging information. Holographic imaging approach based on measuring the refractive index distributions of cells, however, circumvent these problems and offer quantitative and label-free imaging capability. Here, we demonstrate in vivo two- and three-dimensional holographic imaging of circulating blood cells in intact microcapillaries of live mice. The measured refractive index distributions of blood cells provide morphological and biochemical properties including three-dimensional cell shape, haemoglobin concentration, and haemoglobin contents at the individual cell level. With the present method, alterations in blood flow dynamics in live healthy and sepsis-model mouse were also investigated. | physics.bio-ph | physics | Holographic intravital microscopy for 2-D and 3-D imaging intact circulating blood cells in
microcapillaries of live mice
Kyoohyun Kim1,+, Kibaek Choe2,+, Inwon Park3, Pilhan Kim2,* and YongKeun Park1,4,*
1Department of Physics, Korea Advanced Institute of Science and Technology, Daejeon 34141, Republic of
Korea
2Graduate School of Nanoscience and Technology, Korea Advanced Institute of Science and Technology,
Daejeon 34141, Republic of Korea
3Graduate School of Medical Science and Engineering, Korea Advanced Institute of Science and Technology,
Daejeon 34141, Republic of Korea
4TomoCube, Inc., Daejeon 34051, Republic of Korea
+These authors contribute equally to this work.
*Correspondence:
Prof. YongKeun Park,
E-mail: [email protected],
Tel: +82-42-350-2514, Fax: +82-42-350-2510
Prof. Pilhan Kim
E-mail: [email protected],
Tel: +82-42-350-1115, Fax: +82-42-350-1110
Intravital microscopy is an essential tool that reveals behaviours of live cells under conditions
close to natural physiological states. So far, although various approaches for imaging cells in
vivo have been proposed, most require the use of labelling and also provide only qualitative
imaging information. Holographic imaging approach based on measuring the refractive index
distributions of cells, however, circumvent these problems and offer quantitative and label-free
imaging capability. Here, we demonstrate in vivo two- and three-dimensional holographic
imaging of circulating blood cells in intact microcapillaries of live mice. The measured refractive
index distributions of blood cells provide morphological and biochemical properties including
three-dimensional cell shape, haemoglobin concentration, and haemoglobin contents at the
individual cell level. With the present method, alterations in blood flow dynamics in live healthy
and sepsis-model mouse were also investigated.
Introduction
Intravital microscopy (IVM) visualizes intact cells and tissues in live animals with high spatial
resolution and has been used as an invaluable tool for studying diverse pathophysiology ranging from
cell biology1 and immunology2, 3 to tumour biology4, 5. Various imaging techniques have been
developed and used. Confocal scanning microscopy and multi-photon microscopy have been
commonly used for imaging living cells and tissues in vivo6. Even though these microscopic
techniques provide the 3-D shapes of cells and their dynamics, they have limited capability to
quantifying physiological information about cells. Moreover, the use of exogenous labelling agents
significantly limits the use of these techniques only to animal studies. In clinics, bright-field intravital
microscopes have been widely used in clinics because they do not require the use of labelling agents.
However, they only provide limited imaging capabilities − two-dimensional (2-D) qualitative imaging
information.
In this work, we claim that holographic approach can circumvent these limitations in conventional
IVM because it can quantitatively measure refractive index (RI), intrinsic optical properties of
biological cells and tissues. Recently, quantitative phase imaging (QPI) techniques have emerged to
yield quantitative information on individual biological cells and tissues in vitro, without using
exogenous labelling agents7. Exploiting interferometry, QPI techniques measure optical phase delay
images of cells, from which various morphological, biochemical, and biomechanical information can
be retrieved. Due to its unique capability, QPI techniques have been extensively used for the study of
cell pathophysiology, including cell division and growth8, 9, the morphology of blood cells10-13,
neuroscience14-16, parasitology17, 18, and tissue imaging19, 20.
Due to label-free and quantitative capability, these holographic imaging approaches in QPI have
opened the door for several biological and medical studies for cells and tissues in vitro. However, it is
difficult to perform in vivo QPI imaging and for IVM applications, because of light scattering in
tissues. Inhomogeneous distributions in tissues cause a significantly large degree of light scattering21,
which scrambles the optical light fields making it difficult to perform holographic imaging in vivo.
Consequently, QPI for in vivo applications has remained unexplored. Despite the many challenges,
there is strong motivation to extend QPI to in vivo imaging to utilize its unique quantitative and label-
free imaging capability, especially if 3-D imaging of individual cells can be measured.
Here, we present intravital QPI imaging of blood cells circulating in the intact microvasculature of
a live mouse mesentery. A Mach-Zehnder interferometric microscope measures the complex optical
fields of the beam transmitted from the mesentery, from which 2-D quantitative phase maps and 3-D
RI tomograms are retrieved. Light scattering effects from adipose tissues surrounding the
microcapillaries are filtered out, using the fact that the time-varying signals are from moving blood
cells whereas stationary signals are from light scattering by adipocytes. Using the present method, we
demonstrate label-free and quantitative 2-D and 3-D imaging of blood cells in vivo. Biochemical and
morphological properties of individual red blood cells (RBCs) are precisely retrieved, from which
fluid dynamics of blood circulation were systematically investigated. Furthermore, we successfully
visualized alterations in blood circulation in the microvasculature caused by sepsis.
Results
Optical setup
2
In order to perform intravital QPI of microvasculature, we used interferometric microscopy with a
method to filter out static scattering signals. A Mach-Zehnder interferometric microscope was used to
measure optical fields of a coherent laser beam passing through tissues (Fig. 1a). A diode-pumped
solid-state laser (λ = 532 nm, 100 mW, Cobolt Co., Sweden) was used as an illumination source. After
spatial filtering, the beam from the laser is divided into two arms by a beam splitter. One arm is used
as a reference beam. The other beam impinges onto a sample with a long-working distance dry
objective lens (LMPLFLN 50×, NA = 0.5, Olympus Inc., Japan) and a tube lens (f = 200 mm). For
tomographic measurements, the angle of illumination is rotated by using a dual-axis scanning
galvanometer (GVS012, Thorlabs, NJ, USA). Mouse mesentery was staged on a sample plane
equipped with a heating plate (Fig. 1b). The diffracted beam from the sample is collected by either a
high numerical aperture (NA) objective lens (UPLSAPO, 100×, oil immersion, NA = 1.4, Olympus
Inc.) or a water-immersion objective lens (UPLSAPO, 60×, water immersion, NA = 1.2, Olympus
Inc.). The beam is further magnified four times with an additional 4-f telescopic system so that total
field-of-view becomes 35.0 µm × 35.0 µm (high NA objective lens) or 58.4 µm × 58.4 µm (water-
immersion objective lens). The 2-D spatial resolution using the high NA objective lens and the water-
immersion objective lens are calculated as 190 nm and 221.6 nm, respectively, which are defined as
λ/(2NA). The beam diffracted from the sample interferes with the reference beam at an image plane,
which generates spatially modulated holograms. The holograms are recorded by a high-speed CMOS
camera (1024 PCI, Photron USA Inc., CA, USA) with a frame rate of 1,000 Hz and the exposure time
of 1/5,000 sec.
Field retrieval method from scattering tissues
From the recorded raw holograms, complex optical field images of a sample, consisting of amplitudes
and phase delays maps, were extracted with a field retrieval algorithm22. The measured complex
optical fields contain signals from scattered light from surrounding tissues as well as light diffracted
from blood cells. The scattered light from surrounding adipose tissues is significant because of high
RI mismatch between adipocytes and surrounding medium; adipose tissues consist of lipid-storing
3
adipocytes which have high RI (n ~ 1.46)23, 24 compared to that of surrounding medium (for
phosphate-buffered saline (PBS) solution, n = 1.337 at λ = 532 nm)25. Thus, individual circulating
blood cells for our interest are difficult to visualize in the measured optical field images (Fig. 1c), and
retrieved phase delays are contributed by both surrounding tissues and circulating cells (Fig. 1d).
To selectively filter out scattered signals from surrounding tissues, we developed a scattering
reduction method which is shown in Figs. 1e−f. The principle of the scattering reduction method is the
optical field subtraction of stationary signals. The method exploits the fact that even though scattering
from surrounding tissue is significant, they are stationary as long as tissues do not move over time
because light scattering is a deterministic process26. Thus, contributions from light scattering from
surrounding tissues can be removed by subtracting individual optical field images by the time-
averaged optical field image. Then, the resultant optical fields contain information about time-varying
objects, which are circulating blood cells in this case (Fig. 1e). To extract the optical phase delays of
RBCs relative to blood plasma, the optical phase delay of blood plasma inside capillaries independent
of the flow of RBCs is subtracted, which is obtained as the minimum phase values of sequential
frames. The optical phase delay of the RBCs circulating in microvasculature is finally obtained (Fig.
1f). For 3-D tomographic imaging, the scattering reduction method is applied for sequential
holograms with the same illumination angle.
Two-dimensional intravital QPI
To demonstrate the capability of the presented intravital QPI technique and the scattering reduction
method, we first measured 2-D quantitative phase images of live mouse RBCs circulating inside
microvasculature in live mouse mesentery, which was prepared in an imaging chamber. Figure 2a and
Supplementary Movie 1 show the 2-D optical phase delay of individual RBCs flowing inside a
capillary over 0.5 seconds. As individual RBCs are passing through the capillary with a smaller
diameter than the size of RBCs, the quantitative phase maps show shape deformation of RBCs inside
the capillary. In addition to measuring 2-D optical phase delay of RBCs, the present technique can
image other blood components including platelets and white blood cells. For instance, small-sized
4
cells with low optical phase delay indicated in the insets in Fig. 2a are believed to be platelets, which
require further investigation. From 2-D optical phase delay of live RBCs, the dry mass of individual
cells, a mass of nonaqueous component inside cells, can be calculated27, 28. We calculated the dry mass
of 106 individual intact RBCs circulating in microcapillaries, and the distribution of measured
haemoglobin (Hb) contents of individual RBCs was 16.25 ± 3.06 pg [standard deviation of the mean]
(Fig. 2b), which is within the normal range of Hb contents of BALB/c mice29, 30.
In addition to the quantitative measurement of the biochemical characteristics of individual RBCs
from each 2-D phase map, time-lapse phase maps of RBCs flowing inside capillaries provide
quantitative information about the fluid dynamics of live RBCs. The velocity of 106 individual intact
RBCs circulating inside 7 capillaries was measured from the trajectory of the centre of mass of
individual RBCs at each frame. Figure 2c shows that the velocity of RBCs flowing inside
microvasculature is within the range of 100 - 700 µm/s, which was reported previously using flow
cytometry31.
The present technique is also capable of measuring quantitative phase delay of a large number of
RBCs as well as individual RBCs shown in Fig. 2d and Supplementary Movie 2. By summing the
total dry mass of RBCs throughout the field-of-view, the total mass of Hb transported by capillaries
can be quantitatively measured, and the temporal change of the total dry mass provides Hb mass flow
rates of the bloodstream inside capillaries. The change of flow rates of the bloodstream is further
investigated in a bifurcating capillary (Fig. 2e and Supplementary Movie 3), from which the temporal
changes of the dry mass of the incoming and outgoing bifurcated capillaries were measured. For
incompressible fluid flow, the flow rate of the incoming fluid is the same as the sum of flow rates of
the bifurcated outgoing fluid. The flow rate of the bloodstream in bifurcated capillaries, however,
shows complicated fluid dynamics because RBCs interact with capillary walls and neighbouring cells,
especially in a small capillary (Fig. 2f).
Quantitative analysis of live blood flow alteration in a sepsis model with intravital QPI
To extend the applicability of intravital QPI for biological studies, we performed quantitative analysis
5
of the change in blood flow rate in a mouse sepsis model. We intravenously injected
lipopolysaccharides (LPS) solution (20 mg/kg), large molecules found in the outer membrane of
gram-negative bacteria, to induce a severe septic condition which is a dysregulated systemic
inflammatory response including increases in proinflammatory cytokines32. We measured time-lapse
quantitative phase images before and after the LPS injection for an hour for every 15 minutes, and the
same experiment was performed with the injection of PBS solution to compare the effect of the LPS
injection on the change in blood flow. As shown in Fig. 3a and Supplementary Movie 4, the elapsed
time for individual RBCs to pass through a microcapillary is maintained before (34 ms) and at 45
minutes after PBS injection (30 ms). However, the elapsed time (379 ms) for individual RBCs to
passing through a microcapillary was much longer at 45 minutes after the LPS injection compared
with that (19 ms) in the same trajectory of the identical capillary before the LPS injection (Fig. 3b and
Supplementary Movie 5). The difference in the effect of PBS and LPS injection on blood flow is more
clearly seen in the space-time phase images in Figs. 3c-d, where the slope of the phase image of
individual RBCs indicates the flow velocity. The flow velocity of RBCs after LPS injection
significantly decreases compared to the velocity before the injection, while the change in the flow
velocity is minimal in the PBS injection.
For further quantitative analysis, relative volume flow rate and Hb mass rate were measured in 4
and 5 mice for the LPS and PBS groups, respectively (Figs. 3e-f). The relative volume flow rate and
Hb mass rate in LPS injected group decreased quickly compared with those in the PBS injected group.
These results indicate insufficient blood supply to peripheral tissue in LPS-induced sepsis. Decreased
blood flow in capillaries is one of the well-known microcirculatory alterations in sepsis which causes
various organ failures by insufficient supply of oxygen. Decreased systemic pressure and local
arteriolar constriction in sepsis cause blood flow reduction in microcapillaries33, 34. In addition, the
blood flow reduction in microcapillaries may also be a result of decreased RBC deformability in
sepsis which has been described by several mechanisms: LPS or inflammatory cytokines-induced
high production of nitric oxide in sepsis causes decreased RBC deformability through increased
intracellular Ca2+ concentrations in RBCs; reduction of intracellular ATP in RBCs in sepsis also
6
causes increased intracellular Ca2+ concentrations in RBCs through decreased energy for Ca2+
membrane pump, and excessive amount of reactive oxygen species (ROS) released by white blood
cells can alter RBC membranes34.
Three-dimensional intravital QPI
The expansion of the spatial dimension of QPI from 2-D to 3-D can reveal further quantitative
information on biological samples. From 3-D QPI employing optical diffraction tomography (ODT)
technique, the present technique measured the 3-D RI distribution of live RBCs in blood flow, which
provides quantitative structural and biochemical information on the samples including protein
concentration, dry mass, cell volume, surface area, and sphericity12, 35.
Figure 4 and Supplementary Movies 6-7 show 3-D RI distribution of individual RBCs circulating
in a capillary. Cross-sectional slices of 3-D RI tomograms of individual RBCs show that RBCs have
parachute shapes in the capillary (Figs. 4a and 4c) while some RBCs also maintain a discoid shape
with a dimple (Fig. 4b). In addition to the 3-D in vivo visualization of RBCs flowing inside the
capillary, the structural and chemical parameters of RBCs such as Hb concentration, Hb contents, and
RBC volume were extracted from 3-D RI tomograms of individual RBCs (Fig. 4 right panel). The cell
volume, Hb contents, and Hb concentration of 13 intact RBCs were measured as 47.23 ± 6.64 fl,
16.80 ± 2.22 pg, and 35.66 ± 2.38 g/dl, respectively, which are within the range of healthy BALB/c
mice29, 30, 36, 37.
The present technique can also visualize label-free the 3-D structure of microvasculature. The
standard deviation map of the time-lapse 3-D RI distribution of RBCs reveals the 3-D morphology of
a microcapillary (Fig. 4e and Supplementary Movie 8), from which the cross-sectional area of the
microcapillary was calculated as 27.76 µm2. The corresponding capillary diameter is 5.94 µm, slightly
smaller than the diameter of RBCs from BALB/c mice (6.6 µm)38, which implies mechanical
deformation of RBCs passing through the microcapillary. To the best of our knowledge, this is the
first time measuring the quantitative morphological and biochemical parameters of intact individual
cells and microcapillaries in vivo.
7
The reconstructed tomograms of RBCs and microcapillary suffer axial shape elongation as a result
of the finite scanning angle of the long-working distance of the objective lens (NA = 0.5). While the
spatial resolution of the present method is enough to distinguish detailed features of live
microvasculature with sizes ranging in the micrometres, it can be further enhanced by using high-NA
objective lenses or implementing a complex deconvolution method39.
Discussion
In this paper, we present intravital QPI of blood cells inside the microvasculature of live mouse
mesentery. The 2-D intravital QPI clearly shows that the presented technique can measure Hb
contents of individual RBCs in vivo without exposing them in vitro. The present technique also
measures complex optical fields of other blood components which are believed to be platelets.
Because RI increment values for most proteins are similar at 0.190 ml/g40, we expect that the present
technique can also measure other cells in live tissues including white blood cells41, platelets, and
circulating tumour cells.
The 3-D intravital QPI technique reconstructed 3-D RI tomograms of capillaries of live mouse
mesentery, from which structural and chemical parameters of individual RBCs were extracted.
Because the RI of biological samples is an intrinsic optical parameter which is sensitive to
morphological structures and chemical compositions, the present technique has a possibility for
diagnosing live biological tissues and cells under various pathophysiological conditions. While RI
distribution itself does not give molecular specificity for identifying specific proteins in biological
samples, measuring RI dispersion of biological cells by spectroscopic phase microscopy can
distinguish 3-D distribution of various types of proteins label-free42-44.
The time-lapse optical phase maps provide quantitative information on fluid dynamics of RBCs as
well as Hb mass flow rates inside microvasculature, which is an important quantitative measure for
oxygen delivery in the circulatory system. From the time-lapse 2-D optical phase maps, the effect of
LPS on blood flow rate was quantitatively investigated by analysing the volumetric flow rate and Hb
mass rate of microvasculature in live mice. While conventional techniques have measured volumetric
8
flow rate of microvasculature in vivo by multiplying the ensemble velocity of a group of RBCs and
estimated the cross-sectional diameter of blood vessels45, 46, it has been difficult to measure the
amount of Hb transported through capillaries. Because the ability for oxygen transport of the
circulatory system depends on the Hb mass rate related to the number of individual RBCs and
intracellular Hb concentration of RBCs, we believe that the present technique opens a new window
for quantitative investigations on fluid dynamics and oxygen transport rates of blood flow in live
animals in various pathophysiological conditions including RBCs in diabetes47, sickle cell
vasooclusion48 and thrombosis49. We expect that the presented technique can be used to investigate
various disease-related pathological changes in blood cells inside microvasculature in vivo.
The present technique successfully filters out light scattering effects in retrieved complex optical
fields which are originated from adipocyte tissues in live mouse mesentery. Light scattering from
adipocyte tissues is significant. For instance, the reduced scattering coefficient of rat subcutaneous
adipose tissue was measured as
'
sµ = 14.3 cm-1 in the previous research50 and the mean free path, ls,
of a photon in the tissue is 69.9 µm which is calculated as ls = (1-g)/
'
sµ, where anisotropy, g, is
typically 0.9 in biological tissues. It implies that the mesenteric adipose tissue in the present research
suffers at least 2−3 light scattering events. For thicker tissues in which multiple light scattering events
can occur26, we expect that the transmission matrixes (TMs)51-53 or optical phase conjugation
approaches54-56 need to be utilized. This may further expand the applications of QPI to intravital
imaging fields including tumour biology, neurovascular, and cardiovascular biology.
Methods
Animal preparation
BALB/c mice aged 13 and 20 weeks were used for 2-D and 3-D imaging respectively. Mice were
anaesthetized by intramuscular injection of a mixture of Zoletil® (10 mg/kg) and xylazine (11 mg/kg).
The level of anaesthesia was continuously monitored during the experiments with a toe pinch and
maintained by intramuscular injection of half the dose of initially injected Zoletil-xylazine mixture
whenever a response was observed. About 8 cm of the ileum was exteriorized and placed in an
9
imaging chamber. The intestinal loop was surrounded by wet gauze, and the mesentery region was
placed on the glass of the chamber. The exteriorized intestine was maintained at 37℃ with a heating
pad and a temperature sensor. To keep the exteriorized intestine moist, warm saline was supplied into
the gauze in the chamber every 20 min. Because the working distance of the objective lens is limited,
mouse mesentery was fixed on a coverslip by adding the PBS solution. Microcapillaries were
embedded in adipose tissues stacked in approximately 2 to 3 layers with the thickness of
approximately 100 µm. For the mouse sepsis model, lipopolysaccharides (LPS, 20 mg/kg, E. coli
055:B5, L2880, Sigma) dissolved in 100 μl PBS was injected into the tail vein of the anaesthetized
mice. Only PBS (100 μl) was injected into the control mice. The LPS solution and PBS were
prewarmed to about 36℃ immediately before injection. Animal care and experimental procedures
were performed under approval from the Animal Care Committee of KAIST (KA2014-01 and
KA2015-03). All the experiments in this study were carried out in accordance with the approved
guidelines.
Dry mass calculation
The dry mass of individual RBCs (i.e., haemoglobin (Hb) content) was calculated from the retrieved
optical phase delay as follows28:
(
σ
x y
,
)
=
λ
2
πα
(
ϕ
x y
,
)
Hb
=
(
σ
∫∫
x y dxdy
,
)
=
λ
2
πα
(
ϕ
∫∫
x y dxdy
,
)
(1)
(2)
, where
,x yσ
(
)
is the dry mass density (pg/µm2); λ is the wavelength of the illumination beam;
(
,x yϕ
)
is the measured phase map; Hb is the Hb content of individual cells, and α is a RI increment
(α = 0.150 ml/g for both oxygenated and deoxygenated Hb at λ = 532 nm25).
Flow rate analysis
The volumetric flow rate and Hb mass rate were obtained from time-lapse phase images of
10
microcapillaries taken for 500 ms with the frame rate of 1,000 Hz. The volumetric flow rate was
calculated by measuring the velocity distribution of microcapillaries in the space-time phase image of
the microcapillaries. In the space-time phase image along the centre line of a microcapillary,
individual RBCs traveling inside the microcapillary are represented as streaks with slopes
corresponding to the velocity of the blood flow57, 58. The averaged flow velocity, v, was determined
from the Radon transform of the space-time phase map as follows:
v
=
tanx
∆
t
∆
,
θ
max
(3)
where ∆x and ∆t is the spatial and temporal resolution of the space-time phase map, respectively, and
θmax is the angle of the RBC streaks where the maximum variance of the Radon transform of the
space-time phase map is found. The volumetric flow rate, Q, was calculated by multiplying the
averaged flow velocity and the cross-sectional area of the capillary by assuming a cylindrical
geometry for the microcapillary as follows:
Q vA v
=
=
(
r
π
2
)
=
v
π
projA
l
2
2
,
(4)
where A and r is the cross-sectional area and the radius of the capillary; Aproj is the projection area of
the capillary at the normal illumination, and l is the length of the centre line of the microcapillary. The
Hb mass rate was calculated by the Hb dry mass in the area between the initial (e.g., p1 and q1) and
final position (e.g., p2 and q2 in Figs. 3a and 3c, respectively) of the centre line of the microcapillary
divided by the elapsed time for an RBC to pass through the centre line. The calculation was repeated
for individual RBCs passing through the centre line as many times as possible to avoid cell-to-cell
variation.
Tomogram reconstruction
3-D RI distribution of samples was reconstructed from optical fields with various incident angles by
employing ODT. A dual-axis galvanomirror circularly scanned 10 incident beams with various
azimuthal angles with a scanning rate of 10 ms/cycle. The 2-D Fourier spectra of the measured optical
fields with various incident angles were mapped into the 3-D Fourier space according to Fourier
11
diffraction theorem35, 59, 60. Due to sparse illumination and limited acceptance angles originating from
the finite numerical aperture of the objective lens, the resultant 3-D Fourier space has missing
information. The missing information was filled by applying an iterative non-negativity constraint
algorithm. The 3-D RI distribution of samples was then obtained by applying 3-D inverse Fourier
transform. The detailed information about the tomogram reconstruction algorithm via ODT is
presented elsewhere35. Optical field retrieval and tomogram reconstruction were performed by
customized codes in MATLAB® R2014b (MathWorks Inc., MA, USA). RI isosurfaces of individual
RBCs and microcapillaries were rendered by commercial software (Tomocube Inc., Korea).
Acknowledgements
This work was supported by KAIST, KAIST Institute for Health Science and Technology, National
Research Foundation (NRF) of Korea (2015R1A3A2066550, 2014K1A3A1A09063027, 2012-
M3C1A1-048860,
2014M3C1A3052537,
2012M3A6A4054261), Korea Health
Industry
Development Institute (HI15C0399) and
the Innopolis foundation (A2015DD126). K. K.
acknowledges support from Global Ph.D. fellowship from NRF.
Author Contributions
K. K. and K. C. performed experiments and analysed the data. I. P. developed the mouse sepsis model.
Y. P. and P. K. conceived and supervised the project. All the authors wrote the manuscript.
Competing Financial Interests
The authors declare no competing financial interests
12
References
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.
Weigert, R., Sramkova, M., Parente, L., Amornphimoltham, P. & Masedunskas, A. Intravital
microscopy: a novel tool to study cell biology in living animals. Histochem. Cell Biol. 133, 481-491
(2010).
Germain, R.N., Miller, M.J., Dustin, M.L. & Nussenzweig, M.C. Dynamic imaging of the immune
system: progress, pitfalls and promise. Nat. Rev. Immunol. 6, 497-507 (2006).
Phan, T.G. & Bullen, A. Practical intravital two-photon microscopy for immunological research: faster,
brighter, deeper. Immunol. Cell Biol. 88, 438-444 (2010).
Jain, R.K., Munn, L.L. & Fukumura, D. Dissecting tumour pathophysiology using intravital
microscopy. Nat. Rev. Cancer 2, 266-276 (2002).
Beerling, E., Ritsma, L., Vrisekoop, N., Derksen, P.W.B. & van Rheenen, J. Intravital microscopy: new
insights into metastasis of tumors. J. Cell Sci. 124, 299-310 (2011).
Kim, P., Puoris'haag, M., Cote, D., Lin, C.P. & Yun, S.H. In vivo confocal and multiphoton
microendoscopy. J. Biomed. Opt. 13, 010501 (2008).
Lee, K. et al. Quantitative Phase Imaging Techniques for the Study of Cell Pathophysiology: From
Principles to Applications. Sensors 13, 4170-4191 (2013).
Mir, M. et al. Optical measurement of cycle-dependent cell growth. Proc. Natl. Acad. Sci. USA 108,
13124-13129 (2011).
Sung, Y.J. et al. Size homeostasis in adherent cells studied by synthetic phase microscopy. Proc. Natl.
Acad. Sci. USA 110, 16687-16692 (2013).
Popescu, G., Ikeda, T., Dasari, R.R. & Feld, M.S. Diffraction phase microscopy for quantifying cell
structure and dynamics. Opt. Lett. 31, 775-777 (2006).
Park, Y. et al. Measurement of red blood cell mechanics during morphological changes. Proc. Natl.
Acad. Sci. USA 107, 6731-6736 (2010).
Kim, Y. et al. Profiling individual human red blood cells using common-path diffraction optical
tomography. Sci. Rep. 4, 6659 (2014).
Miccio, L., Memmolo, P., Merola, F., Netti, P.A. & Ferraro, P. Red blood cell as an adaptive optofluidic
microlens. Nat. Commun. 6, 6502 (2015).
Marquet, P. et al. Digital holographic microscopy: a noninvasive contrast imaging technique allowing
quantitative visualization of living cells with subwavelength axial accuracy. Opt. Lett. 30, 468-470
(2005).
Mir, M. et al. Label-Free Characterization of Emerging Human Neuronal Networks. Sci. Rep. 4, 4434
(2014).
Jourdain, P. et al. L-Lactate protects neurons against excitotoxicity: implication of an ATP-mediated
signaling cascade. Sci. Rep. 6, 21250 (2016).
Park, Y. et al. Refractive index maps and membrane dynamics of human red blood cells parasitized by
Plasmodium falciparum. Proc. Natl. Acad. Sci. USA 105, 13730-13735 (2008).
Kemper, B. et al. Towards 3D modelling and imaging of infection scenarios at the single cell level
using holographic optical tweezers and digital holographic microscopy. J. Biophotonics 6, 260-266
(2013).
Ding, H.F., Nguyen, F., Boppart, S.A. & Popescu, G. Optical properties of tissues quantified by
Fourier-transform light scattering. Opt. Lett. 34, 1372-1374 (2009).
Bettenworth, D. et al. Quantitative Stain-Free and Continuous Multimodal Monitoring of Wound
Healing In Vitro with Digital Holographic Microscopy. PLoS ONE 9, e107317 (2014).
Tuchin, V.V. Tissue optics : light scattering methods and instruments for medical diagnosis, Edn. 2nd.
(SPIE Press, Bellingham, Wash.; 2007).
Cuche, E., Marquet, P. & Depeursinge, C. Spatial filtering for zero-order and twin-image elimination in
digital off-axis holography. Appl. Opt. 39, 4070-4075 (2000).
Bolin, F.P., Preuss, L.E., Taylor, R.C. & Ference, R.J. Refractive-Index of Some Mammalian-Tissues
Using a Fiber Optic Cladding Method. Appl. Opt. 28, 2297-2303 (1989).
Lai, J., Li, Z., Wang, C. & He, A. Experimental measurement of the refractive index of biological
tissues by total internal reflection. Appl. Opt. 44, 1845-1849 (2005).
Zhernovaya, O., Sydoruk, O., Tuchin, V. & Douplik, A. The refractive index of human hemoglobin in
the visible range. Phys. Med. Biol. 56, 4013-4021 (2011).
Yu, H. et al. Recent advances in wavefront shaping techniques for biomedical applications. Curr. Appl.
Phys. 15, 632-641 (2015).
Barer, R. Interference microscopy and mass determination. Nature 169, 366-367 (1952).
Popescu, G. et al. Optical imaging of cell mass and growth dynamics. Am. J. Physiol. Cell Physiol. 295,
C538-C544 (2008).
13
Tutois, S. et al. Erythropoietic protoporphyria in the house mouse. A recessive inherited ferrochelatase
deficiency with anemia, photosensitivity, and liver disease. J. Clin. Invest. 88, 1730-1736 (1991).
Lee, J.K. et al. Evaluation of the potential immunotoxicity of 3-monochloro-1,2-propanediol in Balb/c
mice - I. Effect on antibody forming cell, mitogen-stimulated lymphocyte proliferation, splenic subset,
and natural killer cell activity. Toxicology 204, 1-11 (2004).
Galanzha, E.I., Tuchin, V.V. & Zharov, V.P. Advances in small animal mesentery models for in vivo
flow cytometry, dynamic microscopy, and drug screening. World J. Gastroenterol. 13, 192-218 (2007).
Doi, K., Leelahavanichkul, A., Yuen, P.S.T. & Star, R.A. Animal models of sepsis and sepsis-induced
kidney injury. J. Clin. Invest. 119, 2868-2878 (2009).
Hinshaw, L.B. Sepsis/septic shock: Participation of the microcirculation: An abbreviated review. Crit.
Care Med. 24, 1072-1078 (1996).
Piagnerelli, M., Boudjeltia, K.Z., Vanhaeverbeek, M. & Vincent, J.L. Red blood cell rheology in sepsis.
Intensive Care Med. 29, 1052-1061 (2003).
Kim, K. et al. High-resolution three-dimensional imaging of red blood cells parasitized by Plasmodium
falciparum and in situ hemozoin crystals using optical diffraction tomography. J. Biomed. Opt. 19,
011005 (2014).
Hawley, R.G. et al. Progenitor cell hyperplasia with rare development of myeloid leukemia in
interleukin 11 bone marrow chimeras. J. Exp. Med. 178, 1175-1188 (1993).
Derelanko, M.J. The toxicologist's pocket handbook, Edn. 2nd. (CRC Press, New York; 2008).
Scott, M.D., Murad, K.L., Koumpouras, F., Talbot, M. & Eaton, J.W. Chemical camouflage of
antigenic determinants: Stealth erythrocytes. Proc. Natl. Acad. Sci. USA 94, 7566-7571 (1997).
Cotte, Y. et al. Marker-free phase nanoscopy. Nature Photon. 7, 113-117 (2013).
Zhao, H.Y., Brown, P.H. & Schuckt, P. On the Distribution of Protein Refractive Index Increments.
Biophys. J. 100, 2309-2317 (2011).
Yoon, J. et al. Label-free characterization of white blood cells by measuring 3D refractive index maps.
Biomed. Opt. Express 6, 3865-3875 (2015).
Fu, D. et al. Quantitative dispersion microscopy. Biomed. Opt. Express 1, 347-353 (2010).
Rinehart, M., Zhu, Y.Z. & Wax, A. Quantitative phase spectroscopy. Biomed. Opt. Express 3, 958-965
(2012).
Jung, J.H., Jang, J. & Park, Y. Spectro-refractometry of Individual Microscopic Objects Using Swept-
Source Quantitative Phase Imaging. Anal. Chem. 85, 10519-10525 (2013).
Riva, C.E., Grunwald, J.E., Sinclair, S.H. & Petrig, B.L. Blood Velocity and Volumetric Flow-Rate in
Human Retinal-Vessels. Invest. Ophthalmol. Vis. Sci. 26, 1124-1132 (1985).
Wang, Y.M., Bower, B.A., Izatt, J.A., Tan, O. & Huang, D. In vivo total retinal blood flow
measurement by Fourier domain Doppler optical coherence tomography. J. Biomed. Opt. 12, 041215
(2007).
Brownlee, M. Biochemistry and molecular cell biology of diabetic complications. Nature 414, 813-820
(2001).
Higgins, J.M., Eddington, D.T., Bhatia, S.N. & Mahadevan, L. Sickle cell vasoocclusion and rescue in
a microfluidic device. Proc. Natl. Acad. Sci. USA 104, 20496-20500 (2007).
Hathcock, J.J. Flow effects on coagulation and thrombosis. Arterioscler. Thromb. Vasc. Biol. 26, 1729-
1737 (2006).
Bashkatov, A.N., Genina, E.A., Kochubey, V.I. & Tuchin, V.V. Optical properties of the subcutaneous
adipose tissue in the spectral range 400-2500 nm. Opt. Spectrosc. 99, 836-842 (2005).
Popoff, S. et al. Measuring the transmission matrix in optics: an approach to the study and control of
light propagation in disordered media. Phys. Rev. Lett. 104, 100601 (2010).
Yu, H. et al. Measuring large optical transmission matrices of disordered media. Phys. Rev. Lett. 111,
153902 (2013).
Yoon, J., Lee, K., Park, J. & Park, Y. Measuring optical transmission matrices by wavefront shaping.
Opt. Express 23, 10158-10167 (2015).
Yaqoob, Z., Psaltis, D., Feld, M.S. & Yang, C. Optical phase conjugation for turbidity suppression in
biological samples. Nature Photon. 2, 110-115 (2008).
Hillman, T.R. et al. Digital optical phase conjugation for delivering two-dimensional images through
turbid media. Sci. Rep. 3 (2013).
Lee, K., Lee, J., Park, J.-H., Park, J.-H. & Park, Y. One-wave optical phase conjugation mirror by
actively coupling arbitrary light fields into a single-mode reflector. Phys. Rev. Lett. 115, 153902 (2015).
Drew, P.J., Blinder, P., Cauwenberghs, G., Shih, A.Y. & Kleinfeld, D. Rapid determination of particle
velocity from space-time images using the Radon transform. J. Comput. Neurosci. 29, 5-11 (2010).
Chhatbar, P.Y. & Kara, P. Improved blood velocity measurements with a hybrid image filtering and
14
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.
iterative Radon transform algorithm. Front. Neurosci. 7, 106 (2013).
Wolf, E. Three-dimensional structure determination of semi-transparent objects from holographic data.
Opt. Commun. 1, 153-156 (1969).
Lauer, V. New approach to optical diffraction tomography yielding a vector equation of diffraction
tomography and a novel tomographic microscope. J. Microsc. 205, 165-176 (2002).
15
Figures with Figure Legends
Figure 1 Schematic diagram of intravital quantitative phase microscopy. a, Mach-Zehnder
interferometry recording holograms from various illumination angles scanned by a galvanomirror
(GM). BS: beam splitter, and M: mirror. b, The schematic diagram of live mouse mesentery staged on
a heating plate. c, Raw holograms recorded by a camera. d, Retrieved optical phase delay maps from
measured holograms in c. e, Optical phase delay maps subtracted by the first frame to eliminate
surrounding adipocytes. f, Final optical phase delay maps subtracted by the minimum values of
sequential phase maps which correspond to the phase delay of red blood cells in the microvasculature
of the mouse mesentery. Scale bar corresponds to 10 µm.
16
a
t = 0.04 sec
t =0.17 sec
t =0.31 sec
(pg/µm2)
1.8
14.02 pg
1.2
13.79 pg
22.64 pg
14.16 pg
t =0.52 sec
t =0.67 sec
t =0.82 sec
0.2
0
0.2
0
13.00 pg
18.00 pg
15.97 pg
19.92 pg
14.54 pg
d
t =0.300 sec
e
t =0.600 sec
ROI 3
ROI 1
ROI 2
161.59 pg
0.6
0
(pg/µm2)
2
1
b
s
C
B
R
f
o
r
e
b
m
u
N
16
12
8
4
0
f
)
g
p
(
s
s
a
m
y
r
D
60
40
20
0
c
30
25
20
s
C
B
R
f
o
r
e
b
m
u
N
15
10
10
Hemoglobin contents (pg)
20
30
5
0
0
100
200
400
300
Velocity (µm/s)
500
600
700
ROI 1
ROI 2
ROI 3
Figure 2 Two-dimensional intravital quantitative phase imaging of individual RBCs in
Time (s)
0
0
0.2
0.4
0.6
0.8
1
microcapillaries. a, Sequential quantitative phase images of individual RBCs flowing through the
microvasculature of the mouse mesentery. Dry mass of individual RBCs is measured from each phase
image. Insets are enlarged view of quantitative phase images indicated by dashed boxes, showing
blood components other than RBCs. Colorbar indicates dry mass densities, and scale bar corresponds
to 10 µm. See also Supplementary Movie 1. b, Histogram of the distribution of measured
haemoglobin contents of individual RBCs in microcapillaries measured from time-lapse phase maps
in a. The mean value of the RBCs haemoglobin contents is 16.25 ± 3.06 pg. c, Histogram of the
velocity distribution of individual RBCs in microcapillaries measured from time-lapse phase maps in
a. d, A snapshot of time-lapse quantitative phase images of the bloodstream flowing inside a capillary.
Total dry mass of the field of view (FOV) is measured from the phase image. See also Supplementary
Movie 2. e, A snapshot of time-lapse quantitative phase images of the bloodstream flowing inside
bifurcating capillaries. Colorbar indicates dry mass densities, and scale bar corresponds to 10 µm. See
also Supplementary Movie 3. f, The change in total dry mass in the regions of interest (ROI) 1, 2 and
3 indicated as boxes in e.
17
a
S
B
P
-
S
B
P
+
b
S
P
L
-
S
P
L
+
t = 103 ms
t = 147 ms
p2
p1
t = 115 ms
t = 145 ms
c
p2
m
µ
5
50 ms
p1
p2
t = 46 ms
t = 65 ms
q1
q2
t = 1 ms
t = 380 ms
m
µ
5
50 ms
p1
d
q2
m
µ
5
q1
q2
50 ms
m
µ
5
50 ms
q1
e
(pg/µm2)
2
1
0
(pg/µm2)
2
1
0
Elapsed Time
Elapsed Time
Before Injection
PBS Injection
LPS Injection
4
3
2
1
0
Q
Q
/
,
l
t
e
a
R
w
o
F
e
m
u
o
V
e
v
i
t
l
l
a
e
R
0
f
0
/
ṁ
ṁ
,
t
e
a
R
s
s
a
M
b
H
e
v
i
t
l
a
e
R
4
3
2
1
0
-5
Injection
15
30
Elapsed Time (min)
Before Injection
45
60
PBS Injection
LPS Injection
-5
Injection
15
30
45
60
Elapsed Time (min)
Figure 3 Quantitative analysis of the effect of LPS injection on live microvasculature. a-b,
Quantitative phase images of individual RBCs flowing through the microvasculature of the mouse
mesentery before (top) and after (bottom panel) the injection of phosphate-buffered saline (PBS)
solution (a) and lipopolysaccharide (LPS) solution (b). Green and blue triangles indicate the initial
and final position of RBCs, respectively, along the microcapillaries for which center lines are
indicated by solid black lines. Scale bars indicate 10 µm. See also Supplementary Movies 4 and 5. c-d,
Space-time phase images of RBCs before (top) and after (bottom panel) the injection of PBS solution
(c) and LPS solution (d), respectively. Individual RBCs indicated as green and blue triangles in the
phase images in a-b are marked as the same triangles. e-f, Changes in relative volume flow rate (e)
and haemoglobin (Hb) mass rate (f) over time before and after the injection of PBS solution (n = 5,
grey boxes) and LPS solution (n = 4, red boxes). Values from individual microcapillaries are
normalized by the value measured before the injection of the solution.
18
Figure 4 Three-dimensional intravital tomographic imaging of individual RBCs in the
microvasculature of the mouse mesentery. a-c, (left panel) x-y cross-sectional slices of the three-
dimensional RI distribution of individual RBCs flowing through a capillary. See also Supplementary
Movie 6. (centre panels) x-y, y-z and x-z cross-sectional slice images of the RI distribution of RBCs
indicated by black arrows in the left panel. (right panel) Rendered isosurfaces of the 3-D RI
distributions of RBCs and measured volume, haemoglobin concentration and haemoglobin contents. d,
19
Rendered isosurfaces of the 3-D RI distributions of RBCs in the microcapillary in c with various
viewing angles. See also Supplementary Movie 7. e, Rendered isosurfaces of the microcapillary with
various viewing angles. See also Supplementary Movie 8. Scale bars indicate 5 µm.
20
|
1206.6562 | 1 | 1206 | 2012-06-28T04:35:58 | From Brownian Dynamics to Markov Chain: an Ion Channel Example | [
"physics.bio-ph"
] | A discrete rate theory for general multi-ion channels is presented, in which the continuous dynamics of ion diffusion is reduced to transitions between Markovian discrete states. In an open channel, the ion permeation process involves three types of events: an ion entering the channel, an ion escaping from the channel, or an ion hopping between different energy minima in the channel. The continuous dynamics leads to a hierarchy of Fokker-Planck equations, indexed by channel occupancy. From these the mean escape times and splitting probabilities (denoting from which side an ion has escaped) can be calculated. By equating these with the corresponding expressions from the Markov model the Markovian transition rates can be determined. The theory is illustrated with a two-ion one-well channel. The stationary probability of states is compared with that from both Brownian dynamics simulation and the hierarchical Fokker-Planck equations. The conductivity of the channel is also studied, and the optimal geometry maximizing ion flux is computed. | physics.bio-ph | physics |
FROM BROWNIAN DYNAMICS TO MARKOV CHAIN:
AN ION CHANNEL EXAMPLE
WAN CHEN1, RADEK ERBAN1,2 , AND S. JONATHAN CHAPMAN1,3
Abstract. A discrete rate theory for general multi-ion channels is presented, in which the continuous dynamics
of ion diffusion is reduced to transitions between Markovian discrete states. In an open channel, the ion perme-
ation process involves three types of events: an ion entering the channel, an ion escaping from the channel, or an
ion hopping between different energy minima in the channel. The continuous dynamics leads to a hierarchy of
Fokker-Planck equations, indexed by channel occupancy. From these the mean escape times and splitting probabil-
ities (denoting from which side an ion has escaped) can be calculated. By equating these with the corresponding
expressions from the Markov model the Markovian transition rates can be determined. The theory is illustrated with
a two-ion one-well channel. The stationary probability of states is compared with that from both Brownian dynamics
simulation and the hierarchical Fokker-Planck equations. The conductivity of the channel is also studied, and the
optimal geometry maximizing ion flux is computed.
Key words. ion hopping, hierarchical Fokker-Planck equations, transition rates, optimal flux
1. Introduction. The membrane of a eukaryotic cell is mainly composed of a lipid
bilayer, which is impermeable to water solvated ions [2]. Ion channels are nanopores formed
by transmembrane proteins; they allow ions to flow through and act as biological valves
connecting the intracellular with extracellular domains. Ion channels are the main mechanism
by which cells control the intracellular concentration of chemical species, as well as the
potential gradient across the membrane. As such they play important roles in maintaining
various functions of plant, animal and human cells.
There are two main features which distinguish ion channels from other nanoscale porous
media. Firstly, they may be selective, distinguishing between the charge and size of ions; for
example, the potassium K+ channel conducts potassium ions at a rate 104 times faster than it
does sodium ions [10]. Secondly, their conformations may change between open and closed
states in response to an external stimulus such as a voltage gradient, ligand binding, or pH
value.
The molecular structure of many ion channels has been revealed by X-ray crystallogra-
phy in recent decades, which provides insight into their features and function. For example,
the potassium K+ channel is composed of four identical subunits which create a cavity con-
necting the cell interior to a selectivity filter at the outer end of the pore [10]. The narrow
selectivity filter is only 12 A long and about 3 A wide, which forces potassium ions with
Pauling radius 1.33 A to shed their hydrating waters to enter and pass in a single-file fashion.
The oxygen atoms of four carbonyl groups form four rings around the selectivity filter, which
generate local minima called binding sites in the overall energy landscape to coordinate the
dehydrated ions.
Mathematical models for ion channels include molecular dynamics (MD), Brownian dy-
namics (BD) and continuum theory (Poisson-Nernst-Planck equations) in descending order
of resolution [8, 7, 15]. Molecular dynamics provides the most detailed description by mim-
icking the motions and interactions of all atoms (from membrane proteins to free ions and
even individual water molecules) at the molecular level [3, 18, 14]. Since the relaxation of
water molecules happens at the fastest timescale of 1 fs, the time step of an MD simulation
has to be very small, and one needs to evolve a system of thousands of particles up to of the
order of 0.1 ms to observe ion conduction. Such a simulation is obviously computationally
1Mathematical Institute, University of Oxford, 24-29 St Giles', Oxford OX1 3LB, United Kingdom
2E-mail: [email protected]
3E-mail: [email protected]
1
intensive, but much shorter simulations (of the order of 10 ps) can be used to obtain informa-
tion about the local potential energy and the effective diffusion coefficient of ions, which can
then be fed into BD simulations.
Brownian dynamics [22, 9, 16, 6] is a more coarse-grained simulation which replaces
the solvent molecules (water) as a continuum, and represent their influence by a dialectric
constant and stochastic forcing. The fluctuations of membrane proteins are ignored and the
channel is approximated by a solid boundary. Because the dynamics of water and proteins
are no longer included, a relatively long time step can be used, which greatly reduces the
computational cost. In this paper, we focus on this level of resolution, and introduce a discrete
rate theory that is based on observations from BD.
The continuum model [5, 21, 17] calculates the potential energy by a mean-field ap-
proximation of average ion positions, which yields a Poisson equation, and then formulates
a Boltzmann equation (in equilibrium) or a Nernst-Planck equation (in non-equilibrium) for
the ion concentration. These continuum partial differential equations (PDEs) can be solved
efficiently; however the individual ion-ion interaction is missing in this mean-field assump-
tion which then fails to predict some properties (e.g. saturation). Comparisons of BD and
continuum theories in different channel configurations are presented in [9, 16].
Recently several hybrid models combining MD and the theory of stochastic processes
have been proposed, which are able to include molecular details and access long time scales
while keeping computational cost low. One idea is to apply the Eyring rate theory to the ion
permeation process using the potential of mean force (PMF) calculated using MD. This is
based on the assumption that channels have some binding sites, and ions pass through by a
hopping mechanism: an ion fluctuates around a certain site before it obtains enough energy
to overcome the energy barrier and hops into the adjacent vacant site. This ion hopping
mechanism has been revealed by MD in channels with binding sites [3, 4, 14]. In addition,
the single file diffusion constraint imposed by the narrowness of the channel assures that ions
cannot cross each other in the channel. Therefore, the continuous dynamics of ion diffusion
can be represented by transitions between discrete Markovian states.
The Eyring rate theory was originally designed for chemical reactions in the 1930s, with
transition rates proportional to the exponential of the energy barrier and distance between
binding sites [11] (as shown in [8] this overestimates the physical barrier in the ion-crossing
process). A novel theory was proposed recently in [1] for a one-dimensional channel with
sawtooth-like PMF, in which the transition rates are not approximated using the energy barrier
but are obtained as the product of total escape rate from one binding site and the splitting
probability determining the relative chance of landing in each neighbouring site. [1] showed
that an optimal size of binding site maximizes the ionic flux if the applied voltage exceeds a
threshold. They assume the channel is occupied by at most one ion, whereby the resulting
system forms a single Markov chain, and the rates can be solved explicitly. In the multi-ion
channel considered here, ion-ion interactions as well as the higher dimension of the energy
landscape mean that the complexity of the rate theory is greatly increased.
In this paper, we present a general discrete rate theory for a multi-ion channel, and com-
pare it with BD. The ion permeation process involves ion hopping, ion escaping and ion
entering. For the purposes of this work we assume ion entry rates are known and focus on
calculating the other rates in terms of the mean escape time and splitting probability. Be-
cause of the complicated network between states the rates are more intricately related to
these quantities than in the single ion case. Moreover, since analytical solutions for the mean
escape time and splitting probability are not available, these must be determined by solving
the corresponding PDEs numerically. The theory is illustrated by a two-ion channel with one
binding site and two ion sources. We show that, as with the one-ion channel, there exists an
2
I
E
x = −L
x = L
FIG. 2.1. A schematic structure of a channel. x = −L and x = L are the (artificial) left and right bound-
aries connecting large reservoirs of electrolyte in and outside the cell respectively. I and E represent the overall
intracellular and extracellular environments, respectively.
optimal shape for the external potential that allows a maximal flux.
The structure of this paper is as follows. In Section 2, we introduce a general theory for
a multi-ion channel with a maximal capacity of N ions. We first present BD simulations and
formulate an equivalent cascade of hierarchical Fokker-Planck equations for the probability
distribution of ions. An illustrative example of a 2-ion channel is discussed and the probability
distribution from the histogram of BD and the solution of the Fokker-Planck equation are
compared. Next, a discrete rate theory framework is presented in Section 3 and the transition
rates calculated. The 2-ion channel is revisited in this framework, and the result is compared
with that from BD. In Section 4, we apply the theory to study the dependence of channel
conduction on different parameters such as the diffusion coefficient, ion entry rate and depth
of potential wells. In particular, we study the effect of the geometry of the external potential
in Section 5. We conclude by discussing the advantages and limitations of this method and
possible applications and extensions in Section 6.
2. Brownian dynamics. In this section, we present the theoretical framework of BD
simulation. Since we are interested in studying the ion permeation process, which occurs on
a time scale of 10−7 s, and since conformational changes occur on a timescale of 10−3 s, we
assume that channel is always open and does not change its conformation.
Since the channel is very narrow and the ions pass through in single file [14], we will
suppose that the motion is one-dimensional, that is, the centres of the ions will be constrained
to lie along a line. The generalisation to a fully three dimensional channel is algebraically
complicated but conceptually straightforward. Since ions cannot pass each other in one di-
mension, we may neglect the finite size of the ions and model them as point particles with
charge.
We define the maximal capacity of a channel to be N, so that it can hold up to N ions at
one time. We denote the number of binding sites in the channel by M. The parameters N and
M vary among different channels; for example, a germicidal A channel has two binding sites
(M = 2) and single-ion occupation dominates (so that N = 1, or perhaps N = 2 to allow for a
knock-on effect) [1, 19].
At the ends of the channel the pore opens out into the intracellular and extracellular
space. A full model would include (probably continuum) models of these spaces, which
would then be joined (preferably matched in terms of matched asymptotic expansions, but
more likely patched) to the channel model. For our present purposes we need to introduce
(artificial) interfaces (i.e. points) at the left and right ends of the channel such that an ion
3
Cell membraneCell membraneExtracellulardomainIntracellulardomain√
2DdWi,
passing through these interfaces is taken to have left the channel and passed into the external
domains. Without loss of generality, we suppose that the left interface connecting the channel
to the intracellular domain lies at x = −L, and the right interface connecting the channel to
the extracellular domain lies at x = L, as shown in Fig. 2.1. Thus an absorbing boundary
condition is imposed at x = −L and x = L.
In BD simulation of an ion channel the contribution of water molecules to the motion of
a solute ion can be approximated by random collisions and an average frictional force in the
evolution equation of the solute ion [22, 16]. The motion of a system of k ions is given by the
Langevin equation
mi dvi = −γ vidt + f k
i = 1, . . . ,k,
i (x1, . . . ,xk)dt + γ
(2.1)
where xi(t) and vi(t) are the location and velocity of the ith ion respectively. There are three
forces on the right hand side of (2.1). The first term corresponds to the frictional force exerted
on the ion by averaging the effect of water molecules; γ is the frictional drag coefficient,
which depends on the surrounding fluid environment. Here we assume it to be uniform so
that γ is constant. The third term is the stochastic force generated by the random collisions
of water molecules; Wi is a Wiener process and D = kBT /γ is the diffusion coefficient, where
kB is the Boltzmann constant and T is the temperature. The second term f k
i (x1, . . . ,xk) is
the overall electric force on the ith ion, including interactions with all other k− 1 ions in the
channel, fixed charges in the protein, and external field across the membrane. It depends on
the locations of all ions, and can be obtained (along with the diffusion coefficient) from MD
simulation.
Note that a typical value of the diffusion coefficient in aqueous solutions at room temper-
ature is D ∼ 10−3mm2s−1, so that the ratio mi/γ ∼ 10−14s−1. Since we usually take a time
step ∆t > 10−12s in the simulation, the system is in an overdamped limit [8]. We may thus
approximate (2.1) by the overdamped Langevin equation
dxi =
D
kBT
f k
i (x1, . . . ,xk)dt +
√
2DdW i,
i = 1, . . . ,k.
(2.2)
The boundary conditions on (2.2) may be described as
1. When the number of ions in the channel k is less than its capacity N, new ions
are generated at the left (respectively right) end at a rate Hk (respectively Gk). In
principle Hk and Gk depend on the current locations of the k ions in the channel
x1, . . . ,xk as well as the intracellular and extracellular environments I and E .
Since we are in the overdamped limit we cannot simply place the incoming ions at
the ends of the channel: under Brownian motion they would immediately cross the
boundary and leave the channel again. Instead we place them at a position within the
channel given by the positional distribution function h(x;x1, . . . ,xk) (or respectively
g(x;x1, . . . ,xk)). Note that h and g also depend on the positions of the existing ions.
This is necessary since, for example, an ion entering the channel from the left must
lie to the left of x1, while an ion entering from the right must lie to the right of xk.
Thus, at the very least, h depends on x1 while g depends on xk.
The functions h and g should be chosen to make the join with the outer model as
smooth as possible, as in [12, 13]. Here we simply assume that h and g, and the
rates Hk and Gk, are given.
2. When xi(t) < −L or xi(t) > L the ith ion is removed from the channel.
3. If xi(t) > xi+1(t) for some i then xi and xi+1 are switched. This enforces the single-
file nature of the channel by preventing an ion overtaking its neighbour. This condi-
tion is unlikely to occur with ions in a channel due to the strong Coulomb repulsion,
but may be necessary if we are interested in neutral molecules.
4
FIG. 2.2. Hierarchical Fokker-Planck Equations describe the conservation of ions in the channel. For k-ion
occupancy, the transitions to and from (k − 1)-ion and (k + 1)-ion occupancy by ions entering and escaping are
demonstrated, along with internal transitions between states.
2.1. Hierarchical Fokker-Planck equations. We denote by Pk(x1, . . . ,xk,t) the proba-
bility density function for the event that there are k ions in the channel at positions x1, . . . ,xk
at time t. Since the number of ions in the channel may run from zero to the channel capacity
N, we have N + 1 such probability density functions. The probability of no ion in the channel
(i.e. k = 0) is denoted by P0(t), and is independent of the spatial variable. We label the ions
by the order of their locations, such that xi < x j for i < j. Then the stochastic process (2.2) is
equivalent to the following hierarchical system of Fokker-Planck equations:
∂t Pk(x1, . . . ,xk,t) = D∇·(cid:16)
−(cid:16)
(cid:17)
∇Pk(x1, . . . ,xk,t)− Pk(x1, . . . ,xk,t)
Fk(x1, . . . ,xk,t)
(cid:17)
1
kBT
Hk(x1, . . . ,xk) + Gk(x1, . . . ,xk)
Pk(x1, . . . ,xk,t)
+ Hk−1(x2, . . . ,xk)h(x1;x2, . . . ,xk)Pk−1(x2, . . . ,xk,t)
+ Gk−1(x1, . . . ,xk−1)g(xk;x1, . . . ,xk−1)Pk−1(x1, . . . ,xk−1,t)
+ Tk+1(x1, . . . ,xk) + Rk+1(x1, . . . ,xk),
(2.3a)
where ∇ = (∂x1, . . . ,∂xk ), Fk = ( f k
1 , . . . , f k
Tk+1(x2, . . . ,xk+1) = D
Rk+1(x1, . . . ,xk) = −D
k ) ∈ Rk, and
(cid:16)∂ Pk+1
1
(cid:16)∂ Pk+1
kBT
1
kBT
− Pk+1
− Pk+1
∂ xk+1
∂ x1
(cid:17)
(cid:17)
(−L,x2, . . . ,xk+1),
(x1, . . . ,xk,L),
f k+1
1
f k+1
k+1
where k = 0, . . . ,N and we use the convention that P−1 = PN+1 = 0. Since only one ion can
escape or enter at any one time, Pk is coupled only to the neighbouring states Pk−1 and Pk+1.
Note that HN = GN = 0, since no ions can enter when the channel is fully occupied.
The first two terms (i.e. the first line) on the right-hand side of (2.3a) correspond to ion
diffusion and ion drift respectively, where the drift term includes the external potential as well
as ion-ion interactions. The third term corresponds to a new ion entering the k-ion channel
from intracellular or extracellular solution; this term is negative since such an event leads to a
transition from a k-ion channel to a (k +1)-ion channel. The fourth and fifth terms correspond
to a new ion entering a (k− 1)-ion channel from the left and right respectively. The sixth and
seventh terms (i.e. the last line) of (2.3a) correspond to ions leaving a (k + 1)-ion channel
from the left and right respectively.
5
k−1 State L . . .Ion Hoppping(k−1) − ion State 2 State 1 . . .k State L Ion Hopppingk − ion State 1 State 2 State L k+1. . .Ion Hoppping(k+1) − ion State 1 State 2 Ion EscapingIon Escaping . . .. . .Ion EnteringIon Enteringk−1kk−1,,kkkk+1k+1,,RRΗGGΗTTThe boundary conditions on (2.3a) are
along with the no-flux condition on the interface xi = xi+1,
(cid:18)∂ Pk
∂ xi
(cid:19)
lim
xi→xi+1
− Pk
1
kBT
f k
i
= lim
xi→xi+1
Pk(−L,x2, . . . ,xk) = 0,
Pk(x1, . . . ,xk−1,L) = 0,
(2.3b)
(2.3c)
(2.3d)
(cid:19)
− Pk
1
kBT
f k
i+1
(cid:18) ∂ Pk
∂ xi+1
for i = 1, . . . ,k − 1, which ensures that the ions are correctly labelled. The reason we have
had to write this as a limit is that the inter-ion potential tends to infinity as xi → xi+1, while Pk
tends to zero. A local analysis shows that we need Pk to tend to zero faster than (xi+1 − xi)2.
We note the following normalisation condition, which holds at all times t,
Pk(x1, . . . ,xk,t)dx1···dxk = 1,
(2.4)
(cid:90)
N
∑
k=0
Γk
where Γk is the available state space when there are k ions in the channel, namely Γk =
{(x1, . . . ,xk) : x1 < x2 < ··· < xk}. We will usually be interested in the steady state; in that
case we solve the coupled hierarchical Fokker-Planck equations for the stationary probability
distribution (cid:101)Pk = limt→∞ Pk(t) for k = 0,1, . . . ,N.
2.2. An example with N = 2. We exemplify the theory above with a simple channel
that is selective to cations with elementary charge e = 1.6× 10−19 C. The selectivity of this
type of channel is generally caused by negative charged boundary proteins, which decrease
the energy barrier imposed by the narrow structure and assist the permeation of cations. For
example, the oxygen atoms of four carbonyl groups in the selectivity filter of the potassium
channel can be modelled by putting four negative partial charges equally spaced on a ring of
radius d that is perpendicular to the x-axis [9].
We consider the simplest possible example of multi-ion channel with capacity N = 2
and a single binding site M = 1. The binding site is located at the position x = ξ and is
a potential well generated by a ring of fixed partial negative charges a distance d from the
channel axis. By Coulomb's law, the potential energy Φ1(x1) seen by one cation at x1 with
charge e traversing through the channel is
(cid:16)
(cid:112)(x1 − ξ )2 + d2
−keZ
(cid:17)
+Ux1
.
Φ1(x1) =
e
kBT
(2.5a)
where ke the Coulomb force constant and Z the total fixed charge on the ring, and U is the
constant field, which imposes a potential difference 2UL across the channel [−L,L]. This
potential difference is small compared to the potential well, and does not change the shape of
the potential well but merely tilts it by a small angle. The force on the ion due to the potential
is
1 = −kBT
f 1
dΦ1
dx1
.
When there are two cations in the channel, at positions x1 and x2, the overall potential
energy Φ2(x1,x2), including the interaction between the two free ions, is
(cid:16)
(cid:112)(x1 − ξ )2 + d2
−keZ
(cid:112)(x2 − ξ )2 + d2
−keZ
+
6
(cid:17)
x1 − x2 +U(x1 +x2)
kee
+
. (2.5b)
Φ2(x1,x2) =
e
kBT
The forces on the two ions are then
1 = −kBT
f 2
∂Φ2
∂ x1
,
2 = −kBT
f 2
∂Φ2
∂ x2
.
Finally we need to specify the entry rates Hk and Gk and entry distribution functions h and
g. We choose the simplest possible model for the entry distribution function. We suppose
that the ions entering from the left are all placed at a position x− near the left-hand end of the
channel, while ions enterying from the right are placed at a position x+ near the right-hand
end of the channel, that is
h(x) = δ (x− x−),
g(x) = δ (x− x+).
We have to be careful in implementing this condition that we preserve the order of the ions in
the channel. We choose to do this as follows: if we are attempting to place an ion at position
x−, and the position of the existing ion x1 < x−, then we abandon the insertion of the new ion.
A similar procedure is implemented at the right-hand end. In effect this means that the rate
of entry is chosen to be zero whenever the position x1 of the existing ion is such that x1 < x−
or x1 > x+. (An alternative procedure would be to modify the distribution functions h and g
so that h = 0 if x1 < x− and g = 0 if x1 > x+, but this would mean altering them from the
present δ -functions.)
In general the entry rates may be functions of the current ion numbers and locations as
well as the intracellular I and extracellular E environments. However, for this illustrative
example we suppose that they are constant subject to the constraint set out above. Thus we
choose
H0 = λ , G0 = µ, H1 = λΘ(x1 − x−), G1 = µΘ(x+ − x1),
where Θ is the Heaviside function. Recall that H2 = G2 = 0 since the channel is then fully
occupied. To run the Brownian simulation, we set the time step ∆t = 100ns, and the physical
parameters as
L = 1nm, x± = ±0.9nm, ξ = 0nm, d = 0.5nm, D = 1nm2 · ns−1,
T = 298K,
kB = 1.38× 10−23 J· K−1, U = 0V· nm−1, Z = e.
λ = µ = 5ns−1,
(2.6)
We use the nanometer as the unit of length and the nanosecond as the unit of time. We evolve
(2.2) for 2× 109 iterations until a dynamic equilibrium is reached. During the simulation the
number of ions in the channel varies in time as ions enter and leave. We record the number
of ions and their locations at each time step. We find that the proportion of time spent with k
ions in the channel, Jk say, is given by
J0 ≈ 0.0000,
J2 ≈ 0.1014.
J1 ≈ 0.8986,
(2.7)
Thus, for these parameters, the channel is almost never empty, and for nearly 90% of the
time there is just one ion in the channel, with two ions the remaining 10% of the time. The
histograms of 2-ion distribution and 1-ion distribution are plotted in Fig. 2.3(a) and Fig.
2.4(a), respectively.
The stationary probability distributions (cid:101)P2(x1,x2), (cid:101)P1(x1) and (cid:101)P0 satisfy the stationary
7
(a) Histogram
(b) (cid:101)P2(x1,x2)
FIG. 2.3. Using the parameters in (2.6), the stationary probability density of a 2-ion channel is computed as
(a) histogram from Brownian dynamics simulation; (b) solution of (2.8a)-(2.8e). Here x1 is the position of the first
ion and x2 is the position of the second ion; since we label the ions such that x1 < x2 the state space is a triangle.
Fokker-Planck equations for a two-ion channel, namely
0 = D
0 = D∇·(cid:16)
(cid:17)
∇(cid:101)P2(x1,x2) +(cid:101)P2(x1,x2)∇Φ2(x1,x2)
+ λΘ(x2 − x−)δ (x1 − x−)(cid:101)P1(x2) + µΘ(x+ − x1)δ (x2 − x+)(cid:101)P1(x1),
(cid:33)
(cid:32)
d(cid:101)P1
+ λ δ (x1 − x−)(cid:101)P0 + µ δ (x1 − x+)(cid:101)P0
(cid:32)
∂(cid:101)P2
(cid:33)
(x1) +(cid:101)P1(x1)
(cid:33)
+(cid:101)P2
d(cid:101)P1
0 = −(λ + µ)(cid:101)P0 + D
− (λΘ(x1 − x−) + µΘ(x+ − x1))(cid:101)P1(x1)
(cid:32)
∂(cid:101)P2
+(cid:101)P2
(cid:32)
d(cid:101)P1
(cid:33)
+(cid:101)P1
(−L,x1)− D
(−L)− D
+(cid:101)P1
∂Φ2
∂ x1
(cid:32)
∂Φ2
∂ x2
dΦ1
dx1
(x1,L),
d
dx1
+ D
∂ x1
(cid:33)
(x1)
∂ x2
dx1
dΦ1
dx1
dx1
(2.8a)
(2.8b)
(cid:101)P2(−L,x2) =(cid:101)P2(x1,L) = 0,
(cid:33)
(cid:32)
∂(cid:101)P2
∂ x1
+(cid:101)P2
∂Φ2
∂ x1
lim
x1→x2
= lim
x1→x2
dΦ1
dx1
dx1
(cid:101)P1(−L) =(cid:101)P1(L) = 0.
(cid:33)
(cid:32)
∂(cid:101)P2
+(cid:101)P2
.
∂Φ2
∂ x2
∂ x2
(L).
(2.8c)
(2.8d)
(2.8e)
with the boundary conditions
and
We solve (2.8a)-(2.8e) by the finite element PDE solver Comsol with 28800 elements.
2.4(b). We see that these agree with the histograms in Fig. 2.3(a) and Fig. 2.4(a) obtained
from Brownian dynamics simulations.
The stationary distribution (cid:101)P2(x1,x2) is shown in Fig. 2.3(b), and (cid:101)P1(x1) is shown in Fig.
We see that (cid:101)P2(x1,x2) is localised around two discrete states near (x−,ξ ) and (ξ ,x+),
while (cid:101)P1(x1) is localised around x1 = ξ . The most likely path between the two states of
(cid:101)P2(x1,x2) can also be faintly seen.
This localisation of(cid:101)P2 and(cid:101)P1 motivates the definition of a small number of discrete states
which the system can adopt, which is the basis for the discrete transition rate theory described
in the next section.
8
(a) Histogram
(b) (cid:101)P1(x1)
FIG. 2.4. Using the parameters in (2.6), the stationary probability density of the one-ion state is computed by
(a) histogram from Brownian dynamics simulation; (b) solution of (2.8a)-(2.8e). Here x1 is the position of the single
ion in the channel.
3. Discrete transition rate theory. We saw in our two-ion example (Figs. 2.3 and
2.4) that (cid:101)P2 was mainly localised around two regions in state space, while (cid:101)P1 was mainly
stationary probability distribution (cid:101)Pk(x1, . . . ,xk) is mainly localised around Lk small regions.
localised around one region. Suppose, in general, that when there are k ions in the channel the
Let us denote these regions by
for i = 1, . . . ,Lk;
then (cid:101)Pk is very small outside ∪Lk
(cid:90)
The idea of discrete rate theory is to replace the continuous variable (cid:101)Pk with a set of discrete
k ⊂ Γk
S(i)
k , so that
(cid:101)Pk(x1, . . . ,xk)dx1 . . .dxk ≈ 0.
i=1S(i)
Γk\∪Lk
i=1S(i)
k
probabilities corresponding to the states S(i)
is the (stationary) probability that (x1, . . . ,xk) ∈ S(i)
independent of spatial variables. In total there are LΣ = ∑N
sum of the probabilities of all LΣ states is unity according to (2.4), that is,
is just a number: it is
k=0 Lk states in the channel, and the
(cid:90)
(cid:101)P(i)
k =
S(i)
k
k , so that
(cid:101)Pk(x1, . . . ,xk)dx1···dxk
k . Note that (cid:101)P(i)
k
(cid:101)P(i)
k = 1.
N
∑
k=0
Lk∑
i=1
We now imagine a Markov chain in which the channel undergoes transitions from one of
these discrete states to another, with the transition probabilities dependent only on the current
state (i.e. no past history is involved). This Markov chain is illustrated in Fig. 2.2. Such
Markov chains for ion channels have been previously considered for a single ion in a many-
well channel [1]. However, multiple occupancy of the channel leads to a more complicated
transition structure.
9
Since only one ion can enter or leave at once (so that(cid:101)Pk is coupled only to(cid:101)Pk−1 and(cid:101)Pk+1)
k may have transitions to and from only the states S(·)
we see that S(i)
general master equation for the time-dependent probability P(i)
k+1(t) +∑
(cid:33)
(cid:124)
k (t) = ∑
P(i)
k−1(t) +∑
α (i, j)
k−1 P( j)
β (i,l)
k+1P(l)
d
dt
m
j
γ (i,m)
k
P(m)
k
(t)
(cid:125)
k (t) is of the form
k , S(·)
k−1 and S(·)
k+1. The
γ (m,i)
k
P(i)
k (t)
,
(3.1)
(cid:125)
(cid:32)
(cid:124)
∑
j
−
k +∑
α ( j,i)
l
l
(cid:123)(cid:122)
influx
(cid:123)(cid:122)
k +∑
β (l,i)
m
Outflux
to S(i)
k
k
k
is the transition rate from S( j)
k
is the transition rate from S( j)
k
where α (i, j)
to
S(i)
k−1, and γ (i, j)
k . Thus α describes the influx of a new
ion, β describes the loss of an ion to the intracellular or extracellular environment, and γ
describes a hopping of the ions within the channel. In fact we expect many of these rates to
be zero, since, for example, when we add a new ion to a channel it must occupy either the
left-most or rightmost potential well.
is the transition rate from S( j)
k
k+1, β (i, j)
to S(i)
The entry rates for new ions α (i, j)
k may be determined from Hk and Gk, which for the
and hopping rates
can be computed from the notation of mean escape time and splitting probability, as
present purposes we are assuming are given. The ion escape rates β (i, j)
γ (i, j)
k
described below.
k
To define the mean escape time we set all the influx probabilities to zero. We then
suppose that the channel initially contains k ions located at positions x1, . . . ,xk. We define
the mean escape time τk(x1, . . . ,xk) to be the average time before the channel undergoes a
transition to a (k − 1)-ion configuration, that is, the average time for one ion to leave the
channel. Using the backward-Kolmogorov equation [20] it can be shown that τk satisfies
Then the mean escape time from state S(i)
k
is given by
∆τk − ∇Φk · ∇τk = − 1
D ,
(x1, . . . ,xk) ∈ Γk
τk = 0 if x1 = −L or xk = L.
(cid:82)
(cid:82)
S(i)
k
S(i)
k
τk[S(i)
k ] =
τkPk dx1···dxk
Pk dx1···dxk
.
(3.2a)
(3.2b)
(3.3)
We now determine a similar expression for τk[S(i)
Equating the two expressions will then provide information on the rates β (i, j)
k ] using the discrete transition rate model.
To this end suppose that the channel is initially in the state S(i)
k
k , so that P(i)
k (0) = 1,
P( j)
m (0) = 0 otherwise. As before the influx rates αk are set to be zero. The master equation
then decouples from those for P(·)
k−1 and P(·)
(3.1) for P(·)
k . Given
(cid:90) ∞
this solution we can determine the mean escape time τk[S(i)
k P( j)
(cid:90) ∞
k+1, and we can solve for P(i)
and γ (i, j)
k (t)dt
τk[S(i)
t β (l, j)
k ] as
(3.4)
k ] =
0
.
k
k
.
j
l
∑
∑
∑
∑
l
j
0
10
β (l, j)
k P( j)
k (t)dt
In calculating the mean escape time we have not distinguished between the case that the first
ion leaves from left end into intracellular electrolyte I and the case where the last ion leaves
from right end into extracellular electrolyte E . However, it is important that the discrete state
model gets the ratio of these probabilities correct, since this is what causes a net ionic flux
through the channel. Thus the second piece of information we use to determine the rates
β (i, j)
is the splitting probability ρk(x1, . . . ,xk). This is defined to be the probability
k
of the first ion to exit was x1 from the left-hand side of the channel, under the condition that
an ion-escaping event from a k-ion to a (k−1)-ion channel has occurred, given that the k ions
started in positions (x1, . . . ,xk) initially. The splitting probability function ρk satisfies,
and γ (i, j)
k
for (x1, . . . ,xk) ∈ Γk
ρk = 0 on xk = L.
As with τk, we can now calculate the splitting probability for state S(i)
∆ρk − ∇Φk · ∇ρk = 0
ρk = 1 on x1 = −L,
(cid:82)
k as
ρk[S(i)
k ] =
ρkPk dx1···dxk
Pk dx1···dxk
.
(cid:82)
S(i)
k
S(i)
k
(3.5a)
(3.5b)
(3.6)
To calculate the splitting probability from the Markov chain we need to separate β (l, j)
into
two individual rates representing the case that an ion leaves to the right into the extracellular
domain, and the case than an ion moves to the left into the intracellular domain, that is, we
write
k
Then the probability that an ion escapes to the left given that it escapes is
+ β−(l, j)
k
.
k
k = β +(l, j)
β (l, j)
(cid:90) ∞
ρk[S(i)
k ] =
(cid:90) ∞
0
∑
l
∑
j
∑
β−(l, j)
l
k
k
0
β−(l, j)
∑
j
k (t)dt +∑
P( j)
P( j)
k (t)dt
(cid:90) ∞
∑
j
l
0
.
(3.7)
β +(l, j)
k
P( j)
k (t)dt
through the
Note that, as in the case of the escape time τk, the right-hand side depends on S(i)
k
initial condition on P( j)
k
.
k
and γ (l, j)
By equating (3.3) with (3.4) and (3.6) with (3.7) we have a number of equations to help
determine the unknown rates β (l, j)
. Since only the left-most (respectively right-
most) ion can escape from the left-hand side of the channel (respectively right-hand side),
many of the rates β (l, j)
If we still do not have enough equations to
determine the remaining β (l, j)
, then it will be necessary to determine some of the
transition rates between internal states. Since these do not involve a change in the number of
ions in the channel, they may be determined by standard techniques.
will infact be zero.
and γ (l, j)
k
k
k
k
Note that to determine the net flow of ions through the channel we will also have to
distingiush between ion entry from the left and from the right, that is, we should also split
k = α−(i, j)
α (i, j)
k
+ α +(i, j)
k
.
However, in most cases (at least) one of these rates will be zero, since it is not possible to
have the same transition between two states occuring with an ion entering from either side.
The one case where this is possible is the transition between an empty channel and a one-ion
channel, which occurs in our example below.
11
FIG. 3.1. The transitions between the four different states: the three circles represent the left entry point, the
binding site and the right entry point; a (green) filled circle indicates the presence of an ion.
3.1. Example of two-ion channel. Now we revisit the example of two-ion channel in
Section 2.2 and illustrate the rate theory using the parameters in (2.6).
We have seen that the channel can exist in a 2-ion, 1-ion or 0-ion state. From Fig. 2.4
we see that (cid:101)P1(x1) is localised around the single region x1 = ξ , so that there is only one
metastable state with one ion in the channel. From Fig. 2.3 we see that (cid:101)P2(x1,x2) is localised
around the two states (x−,ξ ) and (ξ ,x+). Thus there are two metastable states with two ions
in the channel. Thus our Markov chain comprises the four states
: {ξ},
: {(ξ ,x+)},
: {(x−,ξ )},
: {}.
(3.8)
S(2)
2
S(1)
1
S(1)
2
S(1)
0
Thus L2 = 2, L1 = 1, L0 = 1 and overall there are LΣ = 4 states for this channel. These states,
and the transitions between them, are illustrated in Fig. 3.1. The circle at center represents
the binding site x = ξ , and two other circles represent the left and right entry positions x = x±.
A (green) filled circle represents a position occupied by an ion. Note that for the transitions
between S(1)
it is important to distinguish between ions entering and leaving from
0
the right and from the left, so that we can calculate the net flow of ions through the channel.
and S(1)
1
Let us first consider the ion entry rates. We find
α +(1,1)
1
= 0, α−(1,1)
1
= λ , α +(2,1)
1
= µ, α−(2,1)
1
= 0, α +(1,1)
0
= µ, α−(1,1)
0
= λ .
to S(1)
to S(2)
Let us now consider the ion leaving rates β (i, j)
Note that the two zero values arise because the transition from S(1)
1
entering from the left, while that from S(1)
1
Note also that α (i, j)
2 occurs via an ion
2 occurs via an ion entering from the right.
2 = 0 for all i, j since with two ions in the channel is already full to capacity.
. In principle we have six of these to
1 must occur via an ion leaving from
1 must occur via an
, β−(1,1)
, β +(1,2)
and γ (2,1)
.
2
,P(1)
,P(1)
1
to S(1)
= 0. This leaves β−(1,1)
to determine. To these we must add the two hopping rates γ (1,2)
2
,P(2)
2
k
determine. However, since the transition from S(1)
2
the left we know that β +(1,1)
ion leaving from the right, so we know that β−(1,2)
and β +(1,1)
1
Denoting the state of the system by the probability vector P = (P(1)
2
master equation governing the evolution of the Markov chain is then
= 0. Similarly the transition from S(2)
2
0 )T , the
to S(1)
2
2
2
1
2
dP
dt = T · P(t),
12
(3.9a)
S1(1)S0(1)γ2βααββαγ2112β211α00S2S2(2)(1) (2,1)(1,2) −(1,1) −(1,1) +(2,1)+(1,2) −(1,1) −(1,1)+(1,1)+(1,1)where the 4× 4 transition matrix T is given by
T =
−β−(1,1)
2
− γ (2,1)
2
γ (1,2)
2
γ (2,1)
2
β−(1,1)
2
−β +(1,2)
2
− γ (1,2)
2
β +(1,2)
2
0
0
α−(1,1)
1
α +(2,1)
1
1
−α−(1,1)
− β +(1,1)
1
β +(1,1)
1
1
− α +(2,1)
− β−(1,1)
+ β−(1,1)
1
1
.
0
0
α +(1,1)
0
+ α−(1,1)
0
−α +(1,1)
0
− α−(1,1)
0
(3.9b)
As expected, the sum of each column of the matrix T is zero (since the system (3.9a) con-
serves probability), so the matrix is rank deficient. The stationary probability(cid:101)P is the eigen-
vector associated with zero eigenvalue of matrix T .
To calculate the mean escape time and splitting probability we set the all entry rates
to zero and solve (3.9a). To emphasize that this is an auxilliary problem and not the true
Markov chain we denote the probability of lying in each state by q(1)
0 (t)
respectively. Then (3.9a) is
2 (t), q(1)
2 (t),q(2)
1 (t),q(1)
dq(2)
2
dq(1)
2
dt = −(cid:16)
dt = −(cid:16)
(cid:16)
2
dq(1)
dt = β−(1,1)
1
dq(1)
β +(1,1)
0
dt =
1
2 + β−(1,1)
γ (2,1)
2
q(1)
2 + γ (1,2)
2
q(2)
2 ,
β +(1,2)
2
+ γ (1,2)
2
q(2)
2 + γ (2,1)
2
q(1)
2 ,
β +(1,1)
1
+ β−(1,1)
1
(cid:17)
(cid:17)
2 −(cid:16)
q(2)
q(1)
1 .
(cid:17)
q(1)
1 ,
(3.10a)
(3.10b)
(3.10c)
(3.10d)
The first two equations decouple. We start by considering the state S(1)
(3.10a) -- (3.10b) subject to the initial conditions q(1)
λ1 + β +(1,2)
γ (2,1)
2
2 (0) = 1 and q(2)
exp(λ1 t)
q(1)
2
q(2)
2
+ γ (1,2)
(cid:32)
(cid:33)
(cid:33)
=
1
2
2
2 , that is, we solve
2 (0) = 0. This gives
λ2 + β +(1,2)
γ (2,1)
2
2
+ γ (1,2)
2
exp(λ2 t),
(3.11)
1
2
(cid:17)
2 + β +(1,2)
q(1)
+ β−(1,1)
(cid:32)
(cid:32)
λ1 − λ2
1
+
λ2 − λ1
(cid:33)
(cid:17)
(cid:17)
dt
2
β−(1,1)
where λ1,λ2 are two eigenvalues satisfying
λ1 + λ2 = −(cid:16)
λ1λ2 = β−(1,1)
(cid:16)
(cid:90) ∞
Using (3.4), the mean escape time τ2[S(1)
2 ] is
(cid:90) ∞
β−(1,1)
β−(1,1)
β +(1,2)
2
τ2[S(1)
2 ] =
t
2
0
2
+ β +(1,2)
2
2 + γ (1,2)
+ γ (2,1)
,
γ (1,2)
2 + β +(1,2)
2
2
+ β−(1,1)
2
γ (2,1)
2
.
2
2
q(1)
2 (t) + β +(1,2)
q(1)
2 (t) + β +(1,2)
q(2)
2 (t)
q(2)
2 (t)dt
2
2 + γ (1,2)
+ γ (2,1)
2 + β−(1,1)
γ (1,2)
2
β +(1,2)
2
2 + β−(1,1)
γ (2,1)
2
2
13
0
=
β +(1,2)
2
,
(3.12a)
β +(1,2)
2
2
0
0
2
2 ] =
2 ] is
ρ2[S(1)
(cid:90) ∞
(cid:90) ∞
β−(1,1)
and, using (3.7), the splitting probability ρ2[S(1)
β−(1,1)
q(1)
2 (t)dt
2
q(1)
2 (t) + β +(1,2)
β−(1,1)
2
β +(1,2)
2
β +(1,2)
2
+ β−(1,1)
2
2
2 (0) = 0 and q(2)
Similarly, by applying the initial conditions q(1)
β−(1,1)
2 + γ (1,2)
+ γ (2,1)
2
2 + β−(1,1)
2 + β−(1,1)
γ (2,1)
γ (1,2)
2
2
β−(1,1)
β +(1,2)
+ β +(1,2)
γ (2,1)
2
2
2
+ β−(1,1)
β +(1,2)
2
q(2)
2 (t)dt
+ β−(1,1)
γ (1,2)
2
2
γ (1,2)
2 + β +(1,2)
1− ρ2[S(2)
β−(1,1)
β−(1,1)
β +(1,2)
2
τ2[S(2)
2 ] =
2 ] =
=
2
2
2
2
.
γ (2,1)
2
(3.12b)
2 (0) = 1, we find
β +(1,2)
2
,
.
(3.12c)
(3.12d)
2
γ (1,2)
2 + β +(1,2)
2
γ (2,1)
2
Finally we have to consider the escape time and splitting probability for state S(1)
initial condition q(1)
is easily solved to give
1 . With the
1 (0) = 1, equation (3.10c) decouples and
2 (0) = q(2)
2 (0) = q(1)
0 (0) = 0, q(1)
(cid:90) ∞
(cid:90) ∞
1 (t)dt
q(1)
1 dt
β−(1,1)
tq(1)
1 ] =
=
0
0
1
β−(1,1)
1
.
+ β +(1,1)
1
τ1[S(1)
ρ1[S(1)
1 ] =
β−(1,1)
1
1
+ β +(1,1)
1
,
(3.12e)
(3.12f)
The mean escape times τ2, τ1 and the splitting probability ρ2, ρ1 can be obtained by solving
(3.2) and (3.5), respectively with k = 2, 1 by Comsol. Fixing U = 0, ξ = 0, the mean escape
time τ2 in triangular domain Γ2 is plotted in Fig. 3.2(a), the splitting probability ρ2 in Γ2
is plotted in Fig. 3.2(b). Since there is no applied field across the channel (U = 0), and the
potential well is at the center, the external potential is symmetric with x = 0, therefore both
functions are symmetric with x1 + x2 = 0. We take the values of functions at the centre of the
different states and obtain
τ2[S(1)
τ2[S(2)
2 ] = τ2(x−,ξ ) = 0.01113,
2 ] = τ2(ξ ,x+) = 0.01113,
τ1[S(1)
1 ] = τ1(ξ ) = 3.9385× 103,
ρ2[S(1)
ρ2[S(2)
2 ] = ρ2(x−,ξ ) = 0.9964,
2 ] = ρ2(ξ ,x+) = 0.0036,
ρ1[S(1)
1 ] = ρ1(ξ ) = 0.5.
2
= β +(1,2)
Equating these expressions to those of (3.12) we find six equations for the six unknown rates.
Solving these we find
β−(1,1)
= 1.2695× 10−4.
Note that by using the mean escape time and the splitting probability we have not had to
estimate the internal hopping rates γ2 from the Fokker-Planck equation, but have been able
to determine them from the auxillary problems we have solved. We will see in Section 5 that
this is especially useful when the internal states are not so well defined.
2 = 0.3361, β−(1,1)
= 89.8283, γ (2,1)
2 = γ (1,2)
= β +(1,1)
1
1
2
14
(a) τ2(x1,x2)
(b) ρ2(x1,x2)
FIG. 3.2. The parameters are given in (2.6). (a) Mean escape time τ2(x1,x2) for 2-ion transiting into 1-ion
state. (b) Splitting probability ρ2(x1,x2) of ion exiting from left side under the condition that an ion escaping event
occurs.
1
Since there is no external potential gradient in this case (U = 0) the rates are symmetric,
and there is no net flux through the channel. However, we can already make some observa-
tions. Firstly, the rates β±(1,1)
are tiny compared to the others. Thus the channel will switch
between single and double occupancy, but will almost never be empty of ions. We will con-
firm this when we consider the equilibrium occupancy of the channel in the next section.
Secondly, the exit rates β−(1,1)
are about 270 times as large as the hopping rates
γ (2,1)
and. This means that, for these values of the parameters, an incoming ion
2
enters and leaves about 270 times before it manages to replace the bound ion in the potential
well at the centre of the channel.
and β +(1,2)
and γ (1,2)
2
2
2
3.2. Stationary probability of each state. Now that we have determined all the transi-
tion rates in Fig. 3.1, the stationary probability for the number of ions in the channel can be
calculated explicitly as
2 =
1 =
(cid:101)P(1)
2 +(cid:101)P(2)
(cid:101)P(1)
(cid:101)P(1)
0 =
α−(1,1)
1
α−(1,1)
0
1
+α +(1,1)
0
τ2[S(1)
2 ] + α +(2,1)
+ α−(1,1)
1
τ2[S(1)
τ2[S(2)
2 ]
2 ] + α +(2,1)
1
1 + 1
τ1[S(1)
1 ]
1
1
τ2[S(1)
2 ] + α +(2,1)
1
1 + 1
τ1[S(1)
1 ]
α−(1,1)
0
1 + 1
τ1(S(1)
1 )
α−(1,1)
0
0
1
+ α−(1,1)
1
+α +(1,1)
1
α−(1,1)
τ1(S(1)
1 )
+ α−(1,1)
1
+α +(1,1)
1
0
0
1
+α +(1,1)
τ2(S(1)
0
≈ 0.1002,
≈ 0.8998,
τ2[S(2)
2 ]
τ2[S(2)
2 ]
2 ) + α +(2,1)
1
≈ 2.285× 10−5.
τ2(S(2)
2 )
(3.13)
We see that these agree well with the probabilities obtained from a Brownian dynamics simu-
lation in (2.7). The same probabilities may be obtained by integrating the solutions of Fokker-
Planck equations over the configuration space, which gives
(cid:90) L
(cid:90) L
−L
−L
I2 ≡
I1 ≡
(cid:90) x2
(cid:101)P2(x1,x2)dx1 dx2 ≈ 0.1001,
(cid:101)P1(x1)dx1 ≈ 0.8991,
−L
I0 ≡(cid:101)P0 ≈ 8.3× 10−4.
15
(3.14)
4. Current-voltage curve. The most important characteristic of an ion channel is its
In this section, we investigate the channel conductance by examining the
conductance.
current-voltage curve for various values of the channel parameters.
The potential difference across a channel is usually around 100 ∼ 200 mV. We therefore
vary the potential gradient U in the range [−0.1,0.1] V nm−1, which gives a voltage drop in
the range [−200,200] mV for a channel of length 2 nm. Following the framework in Section
3, we compute the transition rates corresponding to each given value of U.
By examining the transition network in Fig. 3.1 we see that there are two different paths
which lead to ions moving from the intracellular (left-hand side) to the extracellular (right-
hand side) domain, namely
PATH 1 : S(1)
1
α−(1,1)
1−−−−→ S(1)
2
γ (2,1)
2−−−→ S(2)
2
β +(1,2)
2−−−−→ S(1)
1 ,
PATH 2 : S(1)
1
β−(1,1)
1−−−−→ S(1)
0
α−(1,1)
0−−−−→ S(1)
1 .
Both paths start with a channel with one ion bound at the potential well. In Path 1 another ion
first enters the channel from the left-hand source to produce a two-ion channel. The two ions
then hop to the right so the new ion lies in the potential well at the centre of the channel. The
ion released from this well then exits the channel at the right. Thus in Path 1 we can think of
a new ion coming in and knocking the present ion out the other side. In Path 2 the ion in the
channel first leaves from the right to leave an empty channel, and then a new ion enters from
the left.
By considering the transition rates we can determine the relative importance of each of
these mechanisms. For the parameters in (2.6) the rate β−(1,1)
is tiny and Path 1 dominates
the current. Note that at equilibrium the ion flux entering from the left is balanced by the flux
(cid:101)P(1)
which leaves from right for each path, so that
0 = β−(1,1)
1 = β +(1,2)
(cid:101)P(1)
(cid:101)P(2)
(cid:101)P(1)
α−(1,1)
α−(1,1)
1
1
1
2
2
0
1
,
.
Similarly ions flow from right to left via the paths,
PATH 3 : S(1)
1
α +(2,1)
1−−−−→ S(2)
2
γ (1,2)
2−−−→ S(1)
2
β−(1,1)
2−−−−→ S(1)
1 ,
PATH 4 : S(1)
1
β +(1,1)
1−−−−→ S(1)
0
α +(1,1)
0−−−−→ S(1)
1 ,
and in equilibrium
Combining the current from each path the net current is given by
,
2
2
α +(1,1)
0
(cid:101)P(1)
(cid:101)P(1)
1 = β−(1,1)
(cid:101)P(1)
(cid:101)P(1)
0 + α−(1,1)
1 − β +(1,1)
(cid:101)P(2)
(cid:101)P(1)
2 − α +(1,1)
1 + β +(1,2)
(cid:16)
(cid:16)
α−(1,1)
β−(1,1)
0 = β +(1,1)
(cid:101)P(1)
(cid:101)P(1)
1 − β−(1,1)
(cid:101)P(1)
0 − α +(2,1)
(cid:101)P(1)
(cid:101)P(1)
(cid:101)P(1)
1
1
1
2
1
0
2
2
1
0
1
.
(cid:17)
(cid:17)
,
(4.1)
α +(2,1)
1
I = e
= e
1
Next we study how the conductance of the channel varies with the external potential energy
parameter d/L and the dimensionless entry rates L2λ /D and µL2/D. We plot in Fig. 4.1 the
current-voltage (I-V) curve for different entry rates when d/L = 0.5. For large voltages the
current saturates since it is limited by the entry rate of ions λ and µ; this effect is illustrated
in Fig. 4.1(b). In Fig. 4.2 we plot the I-V curve for various values of d with a fixed entry
rate. The slopes of these curves give the conductance of channel, which grows initially with
increasing voltage until diminishing as saturation sets in. We see that as the dimensionless
entry rate increases, or the potential well gets shallower, the conductance of the channel
increases.
16
(a)
(b)
FIG. 4.1.
I-V curve for L2λ /D = L2µ/D = 1, 5, 10 with d/L = 0.5; all other parameters are as in (2.6). (a)
for small voltages, for which internal transitions in the channel are rate limiting; (b) for larger voltages, at which
saturation occurs due to the finite rate of entry of ions from the bulk.
FIG. 4.2. I-V curve for d/L = 0.3, 0.4, 0.5 with L2λ /D = L2µ/D = 5. All other parameters are as in (2.6).
5. Optimal geometry of potential. We showed that a channel with shallower potential
well has larger conductivity in Fig. 4.2, if all other physical parameters are fixed. However,
the depth of the potential well is not the only factor that determines the ion flux. Abad et al
[1] studied the flux through a channel of capacity N = 1 with symmetric M-shape potential
energy, and showed that there exists a critial ratio σ of the width of potential well over the
length of channel, at which the flux is maximized. This optimal geometry of potential energy
requires the potential well is neither too narrow (σ → 0) nor too wide (σ → 1).
We now perform a similar analysis of the multi-ion channel. For our N = 2 example we
study how the shape of an external potential on the protein boundary affects the conductivity
of the channel. We vary the distance d and carefully choose the external charge as
2−(cid:0)1 + 0.52(cid:1)−0.5
d−1L−(cid:0)1 + d2L−2(cid:1)−0.5 ,
Z = e
so that the depths of the resulting potential wells are all the same (and so that Z = e, when
d/L = 0.5); the variation of the potential well with d is shown in Fig. 5.1(b). The resulting
group of potential wells with applied field U = −0.05 V nm−1 are plotted in Fig. 5.1(a), the
darker curve corresponds to Z = e, d = 0.5L, the dotted curve shows how the potential wells
17
(a) Φ vs. x
(b) Z vs. d
FIG. 5.1. (a) A group of potential energy wells for d = 0.1,0.2,0.3, . . . ,0.9,1.0 nm are plotted by solid curves
respectively from narrow to wide well, the darker curve corresponds to Z = e, d/L = 0.5, and the dotted curve
corresponds the energy drop due to applied field U = −0.05 V nm−1. (b) The external charge Z as a function of d.
tilt with the applied field. Obviously, when d is small, we have a steep potential drop near the
binding site, and a relatively flat energy landscape near the ends of the channel. When d/L
gets large, the potential well tends to a curve that is steeper near the end of the channel and
flatter near the binding site.
Now we fix U = −0.05 V nm−1, λ = µ = 5 ns−1, and plot in Fig. 5.2 the stationary
probability density function (cid:101)P2(x1,x2) obtained by solving stationary Fokker-Planck equa-
tions numerically for various values of d. Recall that since x1 < x2, we only look at the upper
left triangular domain. We observe that when the potential well is very narrow and steep at
the binding site with d = 0.1 nm, the probability distribution of two ions is localized at two
tiny spots around (x−,ξ ) and (ξ ,x+). As d increases until d = 0.5 nm, the two spots grow a
little larger and shift slightly away from (x−,ξ ) and (ξ ,x+). Thus for d ∈ [0.1,0.5] nm we
can justify the use of our four-state rate theory. For d between 0.5 nm and 0.6 nm, a ridge of
high probability distribution emerges that connects previous two spots at its two tails. This
implies that instead of the two ions being trapped at one of two states, the two can wander
back and forth freely between these two states. As d increases even further to d = 0.9 nm, the
ridge shrinks to its center peak, which corresponds to the two ions both sitting in the potential
well. Thus, for larger values of d, we have effectively only one state for the two-ion occupied
channel.
S(1)
2
: {((x− + ξ )/2, (ξ + x+)/2)},
Thus, for large d, we can define a single-chained three-state system
S(1)
0
: {ξ},
which is similar to four-state system in Fig. 3.1, except that S(1)
and γ (2,1)
there is no hopping γ (1,2)
= α−(1,1)
all the entry rates α−(1,1)
escaping rates are easily calculated as
S(1)
1
are combined and
between them. Following the framework in Section 3,
= µ and α +(1,1)
= λ are prescribed, and the
= α +(1,1)
2 = S(2)
2
: {},
2
1
(5.1)
2
0
0
1
β−(1,1)
k
=
ρk(S(1)
k )
τk(S(1)
k )
,
β +(1,1)
k
=
1− ρk(S(1)
k )
τk(S(1)
k )
,
k = 1,2.
In Fig. 5.3(a), we plot the mean escape time τ2 from the left state S(1)
curve) and the right state S(2)
2 = (−0.9,0) (black solid
2 = (0,0.9) (black dash-dotted curve) in four-state formulation,
18
(a) d = 0.1 nm
(b) d = 0.2 nm
(c) d = 0.3 nm
(d) d = 0.4 nm
(e) d = 0.5 nm
(f) d = 0.6 nm
(g) d = 0.7 nm
(h) d = 0.8 nm
FIG. 5.2. Fixing λ = µ = 5 ns−1, U = −0.05 V nm−1, we plot the stationary probability distribution (cid:101)P2(x,y)
for d = 0.1,0.2,0.3, 0.4,0.5,0.6, 0.7,0.8,0.9 nm, which are obtained by solving (2.8) numerically. It shows the
change from two distinct states at (−0.9,0) and (0,0.9) to a ridge of high probability regime to one distinct state
near (−0.45,0.45), with the change of geometry of potential well.
(i) d = 0.9 nm
2 = (−0.45,0.45) (red solid curve) in
and compare them with τ2 from the balanced state S(1)
the three-state formulation (5.1). Since the potential well is broader as d increases, the second
ion (the one which is not trapped in the well) feels its effect more, resulting in an exponential
increase in the mean escape time. In a channel with descending voltage from left to right
(U = −0.05 V nm−1), it takes a longer time for two ions in the left state (−0.9,0) to escape
than two ions in the right state (0,0.9), so the black solid curve is above the black dashed
2 = (−0.45,0.45) in the middle of channel, it
curve. For two ions in the balanced state S(1)
takes an even longer time, so the red solid curve is above the two black curves.
Notice that the ratio τ2(−0.9,0)/τ2(0,0.9) is largest when 0.5 < d < 0.6. Note also that
for small d the three-state formulation is invalid (so τ2(−0.45,0.45) is meaningless) but that
for large d the ratio τ2(−0.45,0.45)/τ2(−0.9,0) is close to 1: for a broad potential the mean
escape time is insensitive to the precise initial position of the two ions.
In Fig. 5.3(b) we plot the left splitting probability ρ2 at the left state S(1)
(black solid curve) and the right splitting probability 1− ρ2 at the right state S(2)
2 = (−0.9,0)
2 = (0,0.9)
19
(a)
(c)
(b)
(d)
FIG. 5.3. Fixing U = −0.05 V nm−1, we numerically solve (3.2) and (3.7) and obtain the mean escape times
and left splitting probabilities at two states S(1)
2 = (0,0.9) for d between 0.1 nm and 1 nm,
which determine the escaping rates and transition rates by (3.12). We compare it with the results of the three-state
simplified model (5.1). (a) The mean escape time τ2 vs. d at different states. (b) The splitting probability ρ2 vs. d
at different states. (c) The escaping rates β (1) = β−(1,1)
. (d) The transition rates γ (1) = γ (2,1)
and γ (2) = γ (1,2)
. Note that those rates only depend on the mean escape time and left splitting probability, and are
independent of entry rates.
2 = (−0.9,0) and S(2)
2
2
2
and β (2) = β +(1,2)
2
(black dash-dotted curve) in four-state formulation. For large d, the external potential has
a large gradient near both ends of the channel (as shown in Fig. 5.1(a)), which pulls the
new ion introduced at either source towards the center of channel, and into balance with the
existing ion at around (−0.45,0.45). So new entering ions at each source are less likely to
leave from the same end of the channel, and both the left splitting probability at left state
(black solid curve) and the right splitting probability at right state (black dash-dotted curve)
decrease monotonically as d increases. In addition, the left-splitting probability ρ2 at the
2 = (−0.45,0.45) in the three-state formulation is plotted by the red solid
balanced state S(1)
curve, which overlaps with ρ2(−0.9,0). As with the escape time, the splitting probability is
insenstive to the initial position of the ions, and is dominated by the effects of the potential.
Next we compare the escape rates β in the two formulations in Fig. 5.3(c). The black
2 = (−0.9,0) escapes from the
solid curve shows the rate at which an ion at left state S(1)
left side, the black dash-dotted curve plots the rate at which an ion at right state S(2)
2 =
(0,0.9) escapes from right side. The left and right escaping rates at the balanced state
20
2
2
to S(2)
2 = (−0.9,0) and S(2)
depicted by solid curve is above γ (1,2)
2 ), so the escape rates β−(1,1)
+ α +(1,1)
We remark that the mean escape time of the ion from a one-ion channel τ1(S(1)
and β +(1,1)
2 = (−0.45,0.45) in three-state formulation (5.1) are plotted by red solid curve and red
S(1)
dash-dotted curve, respectively. All escaping rates drop rapidly as d increases. When d ≥ 0.7,
the potential well is so broad that the two ions are trapped in the channel for a long time. In
Fig. 5.3(d), we plot the transition rates between states S(1)
2 = (0,0.9).
Due to the inclined voltage, the transition from left state S(1)
2 occurs more often than
2
the other way, namely the transition rate γ (2,1)
by the
dash-dotted curve. When d < 0.4, the transition rates are very low (< 10−4 ns−1), so the two
states are very distinct. For large d, the observed single state in Fig. 5.2 can be treated as
average of the two distinct states.
1 = {0}) is
(cid:28) β +(1,2)
O(105) times larger than τ2(S(1)
.
By (3.13) we see that for there to be an appreciable probability of having no ion in the channel
is extremely small compared to (cid:101)P(1)
1 )−1 ∼ α−(1,1)
we need τ1(S(1)
= λ + µ. In our example, we choose the smallest
and
, and is therefore negligible. Thus, for any entry rates λ = O(1), µ = O(1), only one-ion
2 +(cid:101)P(2)
1 +(cid:101)P(1)
0
occupancy (small entry rates) or two-ion occupancy (large entry rates) is observed most of
In Fig. 5.4(a), fixing λ = µ = 5 ns−1, U = −0.05 V nm−1, we compare the stationary
probabilities of two-ion occupancy (solid curves) and one-ion occupancy (dash-dotted curve)
for various values of d using (i) the four-state formulation (3.8) (black curves), (ii) the three-
state formulation (5.1) (red curves), and (iii) by solving the Fokker-Planck equations using
Comsol (discrete markers). When d is large, the stationary probabilities obtained from all
three methods agree with each other, which confirms that the single balanced state in the
three-state formulation can be treated as an average of the two distinct states (with frequent
transitions) in the four-state formulation. Because the mean escape time grows from O(10−2)
ns to O(102) ns with d increasing as shown in Fig. 5.3(a), a constant entry rate λ = µ = 5
ns−1 leads to a transition from initially one-ion dominant to two-ion dominant channel.
entry rates to be λ = µ = 1, so the resulting (cid:101)P(1)
(cid:101)P(1)
time (i.e. (cid:101)P(1)
(cid:28) β−(1,1)
2 ≈ 1).
2
0
1
0
1
2
0
1
After obtaining the rates and probabilities, we can compare the flux through the channel
from the rate theory and the solution of the Fokker-Planck equations. We integrate the right
hand side of (2.8b) with respect to x1 over [−L,L], which yields
where
f2L =
f2R =
f1R − f1L + (µ + λ )I0 − (µ
−L
(cid:90) L
(cid:90) L
(cid:32)
(cid:32)
−L
D
D
(cid:32)
(cid:32)
d(cid:101)P1
d(cid:101)P1
dx1
dx1
∂ x1
∂(cid:101)P2
+(cid:101)P2
∂(cid:101)P2
+(cid:101)P2
+(cid:101)P1
+(cid:101)P1
dΦ1
dx1
dΦ1
dx1
∂ x2
∂Φ2
∂ x1
∂Φ2
∂ x2
(cid:33)
(cid:33)
f1L = D
f1R = D
x−
−L
(cid:90) L
(cid:90) x+
(cid:101)P1 dx1 + λ
(cid:33)
(cid:33)
(cid:101)P1 dx1)− ( f2R − f2L) = 0,
(cid:90) L
(−L,x1)dx1 =
(cid:90) L
(−L,x1)dx1,
∂(cid:101)P2
∂(cid:101)P2
∂ x1
−L
D
(x1,L)dx1 =
D
−L
∂ x2
(x1,L)dx1,
(5.2)
d(cid:101)P1
d(cid:101)P1
dx1
dx1
(−L) = D
(−L),
(L) = D
(L).
In Table 5.1, we show that the eight transitions connected to state S(1)
1
one by one to the eight terms in (5.2).
in Fig. 3.1 correspond
21
Transitions by escaping
Flux from rate theory
β−(1,1)
2−−−−→ S(1)
1
β +(1,2)
2−−−−→ S(1)
1
β +(1,1)
1−−−−→ S(1)
0
β−(1,1)
1−−−−→ S(1)
0
α−(1,1)
1−−−−→ S(1)
2
α +(2,1)
1−−−−→ S(2)
2
α−(1,1)
0−−−−→ S(1)
1
α +(1,1)
0−−−−→ S(1)
1
S(1)
2
S(2)
2
S(1)
1
S(1)
1
S(1)
1
S(1)
1
S(1)
0
S(1)
0
2
1
1
2
(cid:101)P(1)
β−(1,1)
(cid:101)P(2)
2
(cid:101)P(1)
β +(1,2)
2
(cid:101)P(1)
β +(1,1)
1
β−(1,1)
µ(cid:101)P(1)
λ(cid:101)P(1)
µ(cid:101)P(1)
λ(cid:101)P(1)
1
1
0
2
0
TABLE 5.1
Flux from Fokker-Planck
f2L
− f2R
f1L
− f1R
µI1
λ I1
µI0
λ I0
Transitions computer by the four-state model (3.8) and the hierarchical Fokker-Planck equation.
(a)
(b)
FIG. 5.4. Fixing λ = µ = 5 ns−1, U = −0.05 V nm−1. (a) We plot the stationary probabilities of 2-ion and
1-ion by four-state formulation (3.8) (black curves), three-state formulation (5.1) (red curves) and by solving Fokker-
Planck equation with 28800 elements (discrete marker). (b) We plot the current I vs. d by four-state formulation
(3.8) (black solid curve) and by three-state formulation (5.1) (red solid curve), which are compared with ( fL + fR)
in (5.3) from solution of the Fokker-Planck equation (discrete marker).
The left to right flux across the left boundary of the channel is generated by introducing
new ions µI0 + µ(cid:82) x+−L(cid:101)P1 dx1, and the right to left flux across the left boundary of the channel
introducing new ions λ I0 + λ(cid:82) L
is generated by ions leaving the left boundary f1L + f2L. Similarly we have the left to right
flux across the right boundary of the channel is generated by ions leaving the right boundary
− f1R − f2R, and the right to left flux across the right boundary of the channel is generated by
x−(cid:101)P1 dx1. Thus the overall fluxes are
(cid:90) x+
−L
(cid:101)P1 dx1 − f1L − f2L,
fL = µI0 + µ
fR = − f1R − f2R − λ I0 − λ
(5.3)
(cid:90) L
x−
(cid:101)P1 dx1.
In Fig. 5.4(b), we fix λ = µ = 5 ns−1, U = −0.05 V nm−1, and plot the current I vs.
d from the flux ( fL + fR) in (5.3) by discrete open diamands, obtained by solving (2.8). In
22
comparison, the black solid curve depicts the current obtained by applying the four-state rate
theory (3.8), and the red solid curve depicts the current by the three-state rate theory (5.1). We
see that for a broad potential well with 0.6 < d < 1, all three methods reach a good agreement:
the four-state works for large d, even though the stationary probability distribution in Fig.
5.4(a) shows there should be three states for large d. This is because of our use of escape
times and splitting probabilities to determine the transition rates in the model, rather than by
trying to estimate hopping rates directly. However, with small d, the three-state model is not
an accurate description of the Markov process, and the flux quickly becomes inaccurate.
An important observation from Fig. 5.4(b) is that a maximal current is achieved around
d = 0.6 nm, which means there exists an optimal shape of the potential well to conduct ions,
even if the depth of the well remains the same. This result agrees with the argument in [1] for
a single ion channel with piecewise linear potential energy.
We may explain the existence of an optimal flux by looking at the stationary probabilities
(cid:101)P2 and (cid:101)P1 as a function of d in Fig. 5.4(a). When d is small, the potential well is very narrow
(cid:17)
are over 500 ns−1. Thus the channel has only one ion ((cid:101)P1 > 0.8) most of the time, obviously
and steep near the binding site, but relatively flat near the end of channel, so any new ion
introduced would escape very quickly from the same end by diffusion, leaving the old ion in
S+(1,2)
the channel. For example, at d = 0.1 nm, both escaping rates β2
2
the process of ion entering and leaving from the same end does not generate any through flux.
On the other hand, when d is large, the potential well is very broad and flat near the binding
site, two ions can hardly escape from the channel, as shown by the large mean escape time in
Fig. 5.3(a), once a new ion is introduced to the channel, it quickly moves towards the center,
and settles into a balanced state in the well with the other ion; thus the 2-ion state dominates
((cid:101)P2 > 0.9). In this case the flux is small because two ions are trapped in the channel for a long
S−(1,1)
and β2
(cid:17)
(cid:16)
(cid:16)
2
time.
When neither 2-ion or 1-ion occupancy dominates in the channel, so that there are ade-
quate transitions between the 2-ion distinct states and frequent escapes from the 2-ion to the
1-ion state, a large flux is generated. This explains heuristically why an intermediate potential
well has an optimal geometry.
2
2
2
2
, γ (1,2)
, β +(1,2)
and transition rates γ (1,1)
Finally we investigate how the entry rates affect the optimal flux for the family of poten-
tial wells in Fig. 5.1(a) using the four-state Markov chain formulation (3.8). Recall that the
escape rates β−(1,1)
are determined by the potential
through mean escape time and left splitting probability, and thus are independent of the entry
abilities(cid:101)P2 (black) and(cid:101)P1 (red) for λ = 1,5,20,100 ns−1. The intersection points of each pair
rates. We fix the applied field U = −0.05 V nm−1 and plot in Fig. 5.5(a) the stationary prob-
of curves at which(cid:101)P2 =(cid:101)P1 ≈ 0.5 for λ = 1,5,20,100 are at d ≈ 0.665,0.6,0.54,0.41 respec-
tively. This illustrates the fact that when the potential well is narrow and steep at the binding
site (d small), the escaping rates β−(1,1)
are large (shown in Fig. 5.3(c)), so the
entry rates have to increase in order to have equal probabilities of 2-ion and 1-ion occupancy.
In Fig. 5.5(b), we plot the current I(d) for λ = 1,5,20,100 ns−1. As expected, the
current increases as the entry rates increases, but we also find that the value of d at which
the current is optimised shifts; the critical values of d at which optimal flux is achieved are
respectively d ≈ 0.63,0.6,0.59,0.57. Thus the optimal value of d slightly decreases as the
entry rates increases, which shows that the larger escaping rates of tighter potentials require
larger entry rates to optimize the flux.
and β +(1,2)
2
2
6. Conclusion. We have presented a set of hierarchical Fokker-Planck equations de-
scribing ion permeation in multi-ion channels, and reduced these systematically to a discrete
rate theory. The basis of the reduction is the fact that many channels have internal binding
23
(a)
(b)
FIG. 5.5. Fixing U = −0.05 V nm−1. (a) We plot stationary probability (cid:101)P2 (black) and (cid:101)P1 (red) obtained from
at which (cid:101)P2 = (cid:101)P1 ≈ 0.5 for λ = 1,5,20,100 are at d ≈ 0.665,0.6,0.54,0.41 nm, respectively. (b) We plot current
four-state Markov Chain formulation (3.8) for λ = 1,5,20,100 ns−1. The intersection points of each pair of curves
obtained from four-state Markov Chain formulation for µ = 1,5,20,100 ns−1. The critical values of d at which
optimal flux is achieved are respectively d ≈ 0.63,0.6,0.59,0.57 nm.
sites at which ions sit, so that ions transport by hopping between sites on a slow time scale
while oscillating in the binding sites on a fast time scale. Since the fast oscillation is not key
in determining the conduction rate, we can reduce the continuous dynamics to the slow tran-
sition between the discrete states, and thus provide an efficient way to calculate the current
through the channel. A key component of our reduction was the use of exit times and split-
ting probabilities to determine the discrete hopping rates, rather than trying to estimate these
directly using Kramer's theory for example. This means that the predictions of the discrete
model are accurate even when the internal states are not so well defined.
In contrast to traditional Eyring rate theory [11] and the recent study of a one-ion chan-
nel in [1], we have developed a general theory for multi-ion channels, and have shown an
intricate coupling between transition rates, mean escape time and splitting probability, due
to the complexity of the resulting system of Markovian states. The theory is illustrated by a
two-ion channel, which is the most accessible example that includes the multi-ion complexity.
We have investigated how conductivity of the channel depends on the diffusion coefficient,
potential energy landscape, and the ion entry rate. By varying the geometry of the external
potential while keeping the depth fixed, we observed that when the potential well is narrow
and steep at the binding site, the 1-ion state dominates, but when it is not the 2-ion state dom-
inates. In between there is an optimal geometry which maximizes the ion flux by negotiating
between these two extremes and allowing frequent transitions between the 1-ion and 2-ion
states.
Acknowledgements. This publication was based on work supported by Award No KUK-
C1-013-04, made by King Abdullah University of Science and Technology (KAUST). The
research leading to these results has received funding from the European Research Council
under the European Community's Seventh Framework Programme (FP7/2007-2013)/ ERC
grant agreement No. 239870. Radek Erban would also like to thank Somerville College,
University of Oxford, for a Fulford Junior Research Fellowship; Brasenose College, Univer-
sity of Oxford, for a Nicholas Kurti Junior Fellowship; the Royal Society for a University
Research Fellowship; and the Leverhulme Trust for a Philip Leverhulme Prize.
24
REFERENCES
[1] E. ABAD, J. REINGRUBER, AND M. SANSOM, On a novel rate theory for transport in narrow ion channels
and its application to the study of flux optimization via geometric effects, Journal of Chemical Physics,
130 (2009), p. 085101.
[2] B. ALBERTS, A. JOHNSON, J. LEWIS, M. RAFF, K. ROBERTS, AND P. WALTER, Molecular Biology of the
[3] S. BERNECHE AND B. ROUX, Energetics of ion conduction through the K+ channel, Nature, 414 (2001),
Cell, Garland Science, New York, 2007.
pp. 73 -- 77.
[4]
, A microscopic view of ion conduction through the K+ channel, Proceedings of the National Academy
of Sciences USA, 100 (2003), pp. 8644 -- 8648.
[5] D. CHEN, J. LEAR, AND B. EISENBERG, Permeation through an open channel: Poisson-Nernst-Planck
theory of a synthetic ionic channel, Biophysical Journal, 72 (1997), pp. 97 -- 116.
[6] M. CHENG, A. MAMONOV, W. DUKES, AND R. COALSON, Modeling the fast gating mechanism in the
CIC-0 chloride channel, Journal of Physical Chemistry B, 111 (2007), pp. 5956 -- 5965.
[7] K. COOPER, P. GATES, AND R. EISENBERG, Diffusion theory and discrete rate constants in ion permeation,
Journal of Membrane Biology, 106 (1988), pp. 95 -- 105.
[8] K. COOPER, E. JAKOBSSON, AND P. WOLYNES, The theory of ion transport through membrane channels,
Progress in Biophysics and Molecular Biology, 46 (1985), pp. 51 -- 96.
[9] B. CORRY, S. KUYUCAK, AND S. CHUNG, Test of continuum theories as models of ion channels. II. Poisson-
Nernst-Planck theory versus Brownian dynamics, Biophysical Journal, 78 (2000), pp. 2364 -- 2381.
[10] D. DOYLE, J. MORAIS CABRAL, R. A. PFUETZNER, A. KUO, J. GULBIS, S. COHEN, B. CHAIT, AND
R. MACKINNON, The structure of the potassium channel: molecular basis of K+ conduction and selec-
tivity, Science, 280 (1998), pp. 69 -- 77.
[11] H. EYRING, The activated complex and the absolute rate of chemical reactions, Chemical Reviews, 17 (1935),
pp. 65 -- 77.
[12] M.B. FLEGG, S.J. CHAPMAN, AND R. ERBAN, The two-regime method for optimizing stochastic reaction-
diffusion simulations, Journal of the Royal Society Interface, 9 (2012), pp. 859 -- 868.
[13] B. FRANZ, M.B. FLEGG, S.J. CHAPMAN, AND R. ERBAN, Multiscale reaction-diffusion algorithms: PDE-
assisted Brownian dynamics. 23 pages, available as http://arxiv.org/abs/1206.5860, 2012.
[14] M. JENSEN, D. BORHANI, K. LINDORFF-LARSEN, P. MARAGAKIS, V. JOGINI, M. EASTWOOD,
R. DROR, AND D. SHAW, Principles of conduction and hydrophobic gating in K+ channels, Proceedings
of the National Academy of Sciences USA, 107 (2010), pp. 5833 -- 5838.
[15] D. LEVITT, Modelling of ion channels, Journal of General Physiology, 113 (1999), pp. 789 -- 794.
[16] G. MOY, B. CORRY, S. KUYUCAK, AND S. CHUNG, Tests of continuum theories as models of ion channels.
I. Poisson-Boltzmann theory versus Brownian dynamics, Biophysical Journal, 78 (2000), pp. 2349 -- 2363.
[17] B. NADLER, Z. SCHUSS, A. SINGER, AND R. EISENBERG, Ionic diffusion through confined geometries:
from Langevin equations to partial differential equations, Journal of Physics: Condensed Matter, 16
(2004), pp. S2153 -- S2165.
[18] P. PONGPRAYOON, O. BECKSTEIN, C. WEE, AND M. SANSOM, Simulations of anion transport through
OprP reveal the molecular basis for high affinity and selectivity for phosphate, Proceedings of the Na-
tional Academy of Sciences USA, 106 (2009), pp. 21614 -- 21618.
[19] J. PROCOPIO AND O. ANDERSEN, Ion tracer fluxes through gramicidin A modified lipid bilayers, Biophysical
Journal, 25(2 Pt 2) (1979), p. 8a.
[20] S. REDNER, A Guide to First-Passage Processes, Cambridge University Press, 2001.
[21] Z. SCHUSS, B. NADLER, AND R. EISENBERG, Derivation of Poisson and Nernst-Planck equations in a bath
and channel from a molecular model, Physical Review E, 64 (2001), p. 036116.
[22] W. VAN GUNSTEREN AND H. BERENDSEN, Algorithms for Brownian dynamics, Molecular Physics, 45
(1982), pp. 637 -- 647.
25
|
1711.06617 | 1 | 1711 | 2017-11-17T16:40:57 | Planar Optical Nanoantennas Resolve Cholesterol-Dependent Nanoscale Heterogeneities in the Plasma Membrane of Living Cells | [
"physics.bio-ph",
"physics.optics"
] | Optical nanoantennas can efficiently confine light into nanoscopic hotspots, enabling single-molecule detection sensitivity at biological relevant conditions. This innovative approach to breach the diffraction limit offers a versatile platform to investigate the dynamics of individual biomolecules in living cell membranes and their partitioning into cholesterol-dependent lipid nanodomains. Here, we present optical nanoantenna arrays with accessible surface hotspots to study the characteristic diffusion dynamics of phosphoethanolamine (PE) and sphingomyelin (SM) in the plasma membrane of living cells at the nanoscale. Fluorescence burst analysis and fluorescence correlation spectroscopy performed on nanoantennas of different gap sizes show that, unlike PE, SM is transiently trapped in cholesterol-enriched nanodomains of 10 nm diameter with short characteristic times around 100 {\mu}s. The removal of cholesterol led to the free diffusion of SM, consistent with the dispersion of nanodomains. Our results are consistent with the existence of highly transient and fluctuating nanoscale assemblies enriched by cholesterol and sphingolipids in living cell membranes, also known as lipid rafts. Quantitative data on sphingolipids partitioning into lipid rafts is crucial to understand the spatiotemporal heterogeneous organization of transient molecular complexes on the membrane of living cells at the nanoscale. The proposed technique is fully biocompatible and thus provides various opportunities for biophysics and live cell research to reveal details that remain hidden in confocal diffraction-limited measurements. | physics.bio-ph | physics | Planar Optical Nano-Antennas Resolve
Cholesterol-Dependent Nanoscale Heterogeneities in
the Plasma Membrane of Living Cells
Raju Regmi,1,2 Pamina M. Winkler,1 Valentin Flauraud,3 Kyra J. E. Borgman,1
Carlo Manzo,1 Jürgen Brugger,3 Hervé Rigneault,2 Jérôme Wenger,∗,2 and María
F. García-Parajo∗,1,4
1 ICFO-Institut de Ciencies Fotoniques, The Barcelona Institute of Science and Technology,
2 Aix Marseille Univ, CNRS, Centrale Marseille, Institut Fresnel, UMR 7249, Marseille, France,
3 Microsystems Laboratory, Institute of Microengineering, Ecole Polytechnique Fédérale de
08860 Barcelona, Spain,
Lausanne, 1015 Lausanne, Switzerland,
4 ICREA, Pg. Lluís Companys 23, 08010 Barcelona, Spain
E-mail: [email protected]; [email protected]
Abstract
Optical nano-antennas can efficiently confine light into nanoscopic hotspots enabling single-
molecule detection sensitivity at biological relevant conditions. This innovative approach to
breach the diffraction limit offers a versatile platform to investigate the dynamics of individual
biomolecules in living cell membranes and their partitioning into cholesterol-dependent lipid
nanodomains. Here, we present optical nano-antenna arrays with accessible surface hotspots to
study the characteristic diffusion dynamics of phosphoethanolamine (PE) and sphingomyelin
(SM) in the plasma membrane of living cells at the nanoscale. Fluorescence burst analysis
1
and fluorescence correlation spectroscopy performed on nano-antennas of different gap sizes
show that unlike PE, SM is transiently trapped in cholesterol-enriched nanodomains of 10 nm
diameter with short characteristic times around 100 µ s. Removal of cholesterol led to free dif-
fusion of SM, consistent with the dispersion of nanodomains. Our results are consistent with
the existence of highly transient and fluctuating nanoscale assemblies enriched by cholesterol
and sphingolipids in living cell membranes, also known as lipid rafts. Quantitative data on
sphingolipids partitioning into lipid rafts is crucial to understand the spatiotemporal heteroge-
neous organization of transient molecular complexes on the membrane of living cells at the
nanoscale. The proposed technique is fully bio-compatible and thus provides various oppor-
tunities for biophysics and live cell research to reveal details that remain hidden in confocal
diffraction-limited measurements.
Keywords: optical nano-antenna, nanophotonics, fluorescence correlation spectroscopy (FCS),
live cell membrane, lipid rafts
The plasma membrane plays a major role in cell physiology and is thus of fundamental im-
portance to living systems. The spatial organization and diffusion dynamics of its constituents
(lipids and proteins) occurring at the nanoscale largely influence cellular processes such as trans-
membrane signaling, intracellular trafficking and cell adhesion. 1,2 Recent advances in cell biology
have shown that the plasma membrane is significantly more complex than just a continuous fluidic
system. 3–5 It has been postulated that sphingolipids, cholesterol and certain types of proteins can
be enriched into dynamic nanoscale assemblies or nanodomains, also termed lipid rafts. 6–8 Lipid
rafts have been defined as highly dynamic and fluctuating nanoscale assemblies of cholesterol and
sphingolipids that in the presence of lipid- or protein-mediated activation events become stabi-
2
lized to compartmentalize cellular processes. 2,5,9 However, the true nature of these nanodomains
remains debated with many conflicting evidences and predicated domain sizes in the broad range
of 10-200 nm, primarily because of their transient nature and nanoscopic sizes. 8–14
Early investigations on membrane organization were mostly based on fluorescence recovery
after photobleaching (FRAP) 15 and single particle tracking (SPT). 3,16 Both techniques are limited
either in space (with µ m2 probe area in FRAP) or in time (with millisecond temporal resolution
in SPT). Fluorescence correlation spectroscopy (FCS) is a widely adopted alternative for studying
dynamics and biomolecular interactions. 17 FCS determines the average transit time from statistical
averaging over many individual molecule diffusion events. 18 Its high temporal resolution together
with its rather straightforward data analysis makes FCS an attractive tool to probe the spatiotempo-
ral organization of cell membranes. 10,11,19 However, conventional FCS on confocal microscopes is
unable to resolve the nanoscale organization of lipids due to the limited 200 nm spatial resolution
set by diffraction.
Various approaches have been implemented over the past decade to breach the diffraction limit
in FCS, but membrane studies have so far remained above a 40-50 nm detection size. Stimulated
emission depletion microscopy (STED) constrains the excitation spot down to ∼ 30 nm 20 and
has been combined with FCS to explore the nanoscale dynamics occurring in lipid membranes
on living cells. 21–24 An alternative strategy takes advantage of nanophotonic structures to engi-
neer the light intensity distribution at the nanoscale. 25 Some notable designs include zero-mode
waveguides, 26–31 bowtie structures, 32–34 gold nanorods 35 and sub-wavelength tip based NSOM
probes. 36,37 These various approaches allow to confine the illumination light in the range of 50 to
100 nm. Resonant optical nanogap antennas have shown great potential to further constrain the
laser light on a sub-20 nm scale 38 and greatly enhance the light-matter interactions. 39–42 However,
so far the applications of such resonant nanogap antennas have been mostly employed to probe
fluorescent molecules in solutions at high micromolar concentration. Recently, we have developed
a new class of nano-antennas that maximizes access to the antenna hotspot region together with
extreme planarity and biocompatibility. 43 This methodology has been validated using model lipid
3
membranes, 44 underscoring its high potential to investigate the nanoscale architecture of living
cell membranes.
In this work, we combine FCS with these planar optical nanogap antennas to investigate for
the first time the nanoscopic organization of lipid rafts in the plasma membrane of living cells at
a spatial resolution of 10 nm. The antenna design has been specifically developed for FCS with
sub-diffraction spatial resolution. 41 It combines a central nanogap antenna to create the highly
confined electromagnetic hotspot (of dimensions ∼10 and 35 nm) surrounded by a rectangular
cladding to prevent direct excitation of background molecules diffusing away from the central
nanogap. By applying planarization, etch-back and template stripping methods, we have improved
our initial design to produce arrays of nano-antennas with controlled gap sizes, sharp edges and
planar hotspots facing the upper surface of the sample. 43 Using these planar nano-antennas with
gap sizes down to 10 nm, we investigate here the diffusion dynamics of phosphoethanolamine
(PE) and sphingomyelin (SM) on the plasma membrane of living Chinese hamster ovary (CHO)
cells. Compared to earlier works using confocal FCS, 10,11,19 nanoaperture FCS, 26–30 or STED-
FCS, 21–24 our study is the first to breach into the sub-30 nm spatial scale on living cell membranes.
Together with cholesterol depletion experiments, we provide compelling evidence of short-lived
cholesterol-induced ∼ 10 nm nanodomain partitioning in plasma membranes and discuss the im-
pact of these results in the context of lipid rafts.
The planar antenna platform contains multiple gold nano-antenna arrays with nominal gap sizes
of 10 nm and 35 nm on which a circular cell culture well is mounted for live CHO cell culturing.
Figure 1a,b depicts the strategy chosen for the fluorescence live cell experiment conducted on the
nanogap antenna platform. A 640 nm laser light illuminates a single nano-antenna in the sample
plane of an inverted microscope with a high-NA water immersion objective. Throughout this study,
the linear polarization of the laser beam is set parallel to the antenna main axis so as to excite the
nanogap mode. 43 A highly confined nanometric hotspot of illumination light is created on the
surface of the nanogap region which is in direct contact with the adhered plasma membranes of
living CHO cells. Importantly, the planarization strategy avoids possible curvature induced effects
4
on the cell membrane and thus provides an ideal platform for live cell membrane research 29,30
(Supporting Information Fig. S1 shows AFM images indicating a planarity better than 3 nm for the
top surface).
The cells were incubated on the nano-antennas at 37o C for nearly 48 hours prior to the experi-
ments to allow them to freely grow and adhere onto the antenna platform. Lipid analogs (either PE-
or SM-BSA complexes) labeled with the lipophilic organic dye Atto647N were incorporated into
the plasma membrane of the living cells just before the fluorescence measurements (see Methods
for details on staining protocol). The choice of Atto647N as fluorescent dye allows an excellent
overlap with the antenna's main plasmonic resonance (Fig. S2), maximizing the fluorescence en-
hancement in the nanogap. Figure 1c shows a representative confocal image of the morphology
of the CHO cells adhered on a glass coverslip taken after the incorporation of the fluorescent lipid
analogs.
Figure 2a,b shows representative single-molecule fluorescence time traces for PE and SM in
the confocal and in the nano-antenna configuration. The resolution given by the diffraction limited
spot in the confocal scheme does not allow to resolve heterogeneities that may occur at the sub-
200 nm spatial scale, and as a result, the time traces for both PE and SM appear indistinguishable.
In contrast, the highly confined surface hotspot originating from the 10 nm gap antenna clearly
reveals differences in the characteristic diffusion dynamics for PE and SM. As shown in Fig. 2b,
PE displays sharp peaks in the fluorescence time trace as a result of the sub-diffraction excitation
hotspot created by the planar nanogap antenna. Unlike PE, the signature of SM is discernibly dif-
ferent at the nanoscale: the short bursts (a hallmark of free diffusion in ultra-small detection areas)
are accompanied by high intensity bursts of significantly longer durations. This is a direct indi-
cation that the nanoscopic diffusion of SM on the cell membrane is deviating from free Brownian
diffusion as compared to larger macroscopic scales.
To provide more quantitative information about the fluorescence time traces, we performed a
fluorescence burst analysis to represent the distributions of burst duration versus burst intensity
(see Methods). 34 Figure 2c shows the results for both PE and SM for the 10 nm nanogap antenna
5
compared to the confocal configuration. The scatter plots for PE and SM in the confocal configura-
tion show no visible differences with burst durations in the range 1-100 ms and intensities around
20-30 counts/ms. However, in stark contrast, the distributions obtained on the nano-antennas show
clear differences between PE and SM. Diffusion events in sub-ms time scales are notably observed
with the nano-antennas exhibiting burst durations as short as 10 µ s. Such short events are more
than two orders of magnitude faster than in the case of the confocal reference. Regarding the diffu-
sion dynamics for PE (red dots) probed with the nanogap antennas a general trend can be deduced,
namely, brighter events arise at shorter timescales. These can be understood as the detection of a
"best burst event" directly resulting as a consequence of an individual molecule diffusing through
the hotspot in the optimal position and orientation for maximum enhancement. The tighter the
excitation beam confinement, the higher is the local intensity which leads to higher fluorescence
intensity and shorter burst duration (see Supplementary Information Fig. S3 and S4 for additional
fluorescent time traces and analysis on different antennas and cells). We thus relate the events with
burst durations below 1 ms to the trajectories occurring within the nanogap region. 34 In the case of
PE, the bursts with durations above 1 ms feature a lower intensity in the range of 20-70 counts/ms,
which is only slightly increased as compared to the confocal level. We assign these longer burst
duration events to the residual excitation of diffusing molecules within the larger 300 × 140 nm2
box aperture region where the electromagnetic field intensity enhancement is negligible and thus
comparable to the confocal reference.
In contrast to PE, SM probed with the nano-antenna arrays shows a significantly broader dis-
tribution of burst lengths against peak burst intensities (Fig. 2c). High intensities are observed for
burst durations below and above 1 ms. Since these events were not observed for PE, we relate
their occurrence to nanoscopic heterogeneities such as transient molecular complexes on the cell
membrane hindering the diffusion of SM. To support this conclusion, we perturbed the cholesterol
composition in the cell membrane with methyl-β -cyclodextrin (MCD) as cholesterol is expected
to play a significant role in the formation and stability of the lipid nanodomains. The result of the
burst analysis for SM after MCD treatment recovers a distribution which closely resembles the one
6
for PE (Supporting Information Fig. S5). In other words, the intense bursts of duration between
0.1 and 10 ms disappear after cholesterol depletion, consistent with the loss of nanodomains. Alto-
gether, the results from the fluorescence burst analysis demonstrate the benefits of planar nanogap
antennas to explore the nanoscopic organization of lipids in live cell membranes. Clear differ-
ences between PE and SM diffusion dynamics are unveiled that otherwise would remain hidden in
confocal measurements.
To further support these results, we performed fluorescence correlation spectroscopy (FCS)
analysis. FCS records the fluorescence intensity fluctuations as the fluorophores transit through
the detection spot. These fluctuations are analyzed by computing the temporal autocorrelation
function, averaging over thousands of single-molecule diffusion events. We used two different gap
sizes (10 and 35 nm) to quantify the lipid dynamics for increasing detection areas in cell mem-
branes. Figure 3a,b shows the normalized correlation traces for PE and SM in case of the nano-
antennas and the confocal reference. Each of these traces is taken on an individual nano-antenna
(more traces are shown in Fig. S3 and S4 to demonstrate the consistency of our results). Similar
to the burst analysis, we find no significant differences between the FCS curves for PE and SM for
the confocal reference (gray circles in Fig. 3a,b the overlay of the confocal FCS data is shown in
Fig. S6), yielding comparable diffusion times of 25 ± 4 ms (PE) and 30 ± 4 ms (SM), respectively.
In the case of the nano-antennas, we observe that decreasing the gap size leads to a faster diffu-
sion, confirming that the fluorescence signal stems from the nanogap region. We use a two-species
model to fit the FCS data in order to account for the fluorescence contributions stemming from the
nanogap and from the surrounding aperture area (see details in Methods section). A key feature
in FCS is that the molecules contribute to the correlation amplitude in proportion to the square of
their fluorescence brightness, hence the signal from molecules in the nanogap experiencing maxi-
mum enhancement will have a dominating contribution to the FCS curves. 45 The complete results
and values for the FCS fits are detailed in the Supporting Information Tables S1-S3.
The differences between PE and SM diffusion dynamics are highlighted in Fig. 3c where a
direct comparison of the FCS data for the 10 nm gap antenna is shown for different fluorescent
7
lipid analogs. Contrarily to the confocal case (Fig. S6), the difference in diffusion times between
the two lipids becomes more prominent at the nanoscale, with PE exhibiting diffusion times of
0.25 ± 0.06 ms and SM of 0.35 ± 0.04 ms. Moreover, after MCD treatment, the diffusion dynamics
for cholesterol-depleted SM closely resembles that of PE with a diffusion time of 0.19 ± 0.03 ms
(Fig. 3d). These FCS results confirm the presence of cholesterol-enriched nanodomains hindering
the diffusion of SM, in agreement with the results found for the fluorescence burst analysis. In
addition, we retrieved an anomaly value alpha for SM that depended on the probed area, deviating
from unity as the illumination area reduced, from α ∼ 0.85 (for the 35 nm gap antenna) to α ∼ 0.65
(for the 10 nm gap antenna), which is fully consistent with hindered diffusion (see Table S1 to S3).
In contrast, the α values were significantly larger and closer to unity for the cases of PE and SM
after MCD treatment (α ∼ 0.85) and did not depend on the probe area, as expected for Brownian,
unhindered diffusion.
To further analyze and exploit the FCS data we take advantage of the large number of planar
nano-antennas with controlled gaps to carry out a FCS analysis over 60 different antennas and
cells. This approach follows the so-called FCS diffusion law, 19,28 which is a representation of the
diffusion time versus the detection area. Extrapolation of the experimental curve to the intercept
with the time axis provides information on the type of diffusion exhibited by the molecule, i.e.,
free diffusion is characterized by a linear curve crossing the origin (0,0), while hindered diffusion
due to the occurrence of nanodomains leads to a positive intercept on the time axis. 19,28 The nano-
antenna detection area was estimated as the product of the gap size (measured by transmission
electron microscopy TEM) times the full width at half maximum for the intensity profile along
the direction perpendicular to the antenna main axis (simulated by finite difference time domain
(FDTD) method, see Supporting Information Fig. S7 for simulations results). Moreover, the area
sizes were further confirmed by calibration measurements both on freely diffusing dyes in solution
and on pure PE lipid bilayers using antennas of different gap sizes. 44 The 10 and 35 nm gap
antennas correspond respectively to 300 nm2 and 1250 nm2 illumination areas. As the diffusion
time proportionally scales with the detection area, the diffusion coefficient D is retrieved from
8
the slope of the linear fit matching the measured transient diffusion times obtained from the FCS
curves versus the effective detection areas according to the relation D = probe area/4 × τdiff. 18
Figure 4a-c summarizes the characteristic diffusion times for PE, SM and SM after cholesterol
depletion for two antenna gap areas. The extension to include the confocal data is shown in Fig. S8.
From these graphs we derive the following three values plotted in Fig. 4d-f: the diffusion coeffi-
cient (from the slope), the time axis intercept (by extrapolating the linear fit for vanishing probe
area) and the normalized spread in the data points (defined as the width of upper and lower quartiles
divided by the median value). The diffusion coefficients derived from nano-antenna measurements
are DPE = 0.44 ± 0.07 µ m2/s, DSM = 0.38 ± 0.19 µ m2/s and DMCD-SM = 0.46 ± 0.07 µ m2/s
(Fig. 4d) and they are consistent with the confocal measurements and values reported indepen-
dently using STED-FCS. 21 These coefficients represent the diffusion speed in the lipidic region
between the nanodomains, with an additional contribution from diffusion within the nanodomains
and diffusion of the domains themselves.
Extrapolating the fits in Fig. 4a-c towards diminishing probe area leads to the intercepts with
the time axis as summarized in Fig. 4e. The almost zero intercept hitting the origin observed for
PE confirms the expected free Brownian motion diffusion mode. In stark contrast, SM features a
positive y-intercept of about 100 µ s, which highlights a significant deviation from free Brownian
diffusion and the occurrence of nanoscopic domains hindering SM diffusion. Depletion of choles-
terol results on SM diffusion with a close-to-zero time intercept, demonstrating the crucial role
of cholesterol establishing the nanodomains and hindering SM diffusion. Such small nanoscale
heterogeneities have never been detected so far with confocal microscopy, although STED-FCS
down to 1000 nm2 detection area could infer their occurrence. 21 Our results are fully aligned with
these previous findings and importantly, we further reduce the detection areas down to 300 nm2 .
Lastly, we take a closer look at the statistical dispersion of the diffusion times for each gap
area, and introduce the normalized data spread as the width from upper to lower quartiles divided
by the median value (Fig. 4f). The spread in diffusion times for PE and SM after MCD treatment
remains under 25% and can be partially assigned to nanometer variations of the gap size between
9
nanoantennas. 44 These variations stem from the nanofabrication process as a consequence of the
finite grain size of gold and/or the scattering of electrons used during the electron beam lithography.
In contrast to PE and MCD-SM, the data for SM features a significantly higher statistical dispersion
around 50%, which cannot be related solely to dispersion in the nanoantenna sample, but instead
it results from large variations in the SM diffusion behavior, as already noted for the fluorescence
burst analysis (Fig. 2c). These results are fully consistent with the presence of cholesterol-enriched
nanodomains affecting SM diffusion.
Altogether, our results provide compelling evidence for the existence of highly transient and
fluctuating nanoscale assemblies of sterol and sphingolipids in living cell membranes. These exper-
imental observations stand in excellent agreement with the notion that without stabilizing proteins,
lipid rafts can be viewed as intrinsic nanoscale membrane heterogeneities that are small and highly
transient. 2,6–8 We estimate the characteristic residence time of the fluorescent SM lipid analogs in
the nanodomain from the y-intercept in Fig. 4b,e, and find a value around 100 µ s . The typical size
of the nanodomains could in principle also be deduced from the FCS diffusion laws which should
feature a characteristic transition from confined to normal diffusion. 19,28 As we do not observe this
characteristic transition in our data, we conclude that the typical size of the nanodomain is smaller
than the smallest gap size of our nanoantenna, that is 10 nm. Both the typical nanodomain size
about 10 nm and the transient time about 100 µ s stand in good agreement with the predictions
from stochastic models 46 and recent high-speed interferometric scattering (iSCAT) measurements
on mimetic lipid bilayers containing cholesterol. 12 We believe that this shorter characteristic time
as compared to earlier experimental works using STED-FCS 21–23 is related to the smaller 10 nm
resolution achieved in our case. The nanoantenna approach is straightforward to implement on
any confocal microscope equipped for FCS as contrarily to STED, it does not require adding any
supplementary illumination beam. As additional advantage, the excellent planarity of the surface
rules out any artefact potentially induced by the curvature of the cell membrane. 29 We believe that
these advantages and the excellent spatiotemporal resolution largely compensate for the need for
nanofabrication and the more complex FCS fitting procedure.
10
In conclusion, we have demonstrated the promising approach of exploiting planar optical nano-
antennas with accessible surface nanogaps to investigate the nanoscale architecture of live cell
membranes. The key strengths of our approach rely on the 10 nm spatial resolution combined
with a microsecond time resolution on a nearly perfectly flat substrate compatible with live cell
culturing. The single-molecule data on nanoantennas reveal striking differences between PE and
SM diffusion dynamics that remain hidden in confocal measurements. Fluorescence burst and
correlation spectroscopy analysis for PE are consistent with a free Brownian diffusion model. In
contrast, the diffusion dynamics of SM at the nanoscale show heterogeneities in both time and
space which are cholesterol dependent. Indeed, removal of cholesterol leads to a recovery of free
Brownian diffusion for SM, consistent with the loss of nanodomains. Our results are consistent
with the existence of dynamic nanodomains on the plasma membranes of living cells of ∼10 nm
diameter which is comparable to our measurement gap size. The corresponding transient trap-
ping times are short of about ∼100 µ s. We believe that the combination of optical nano-antennas
with fluorescence microscopy has a high potential to investigate the dynamics and interactions of
raft associated proteins and their recruitment into molecular complexes on the plasma membrane
of living cells. The proposed technique is fully bio-compatible and thus provides ample oppor-
tunities for biophysics and live cell research with single-molecule sensitivity at nanometric and
(sub)microsecond spatiotemporal resolution, far beyond the diffraction limit of light.
Methods
Planar Nanogap Antenna Array Fabrication Large scale nano-antenna arrays with surface
nanogaps were fabricated by combining electron beam lithography with planarization, etch back
and template stripping. 43 First, EBL was used to pattern features with negative tone hydrogen
silsesquioxane (HSQ) resist on top of 100 mm silicon wafer. A 50 nm thick gold film was then
deposited by electron beam evaporation at low temperature to reduce the gold grain size. Flowable
oxide (Dow Corning FOX-16) was spun to planarize the overall structure and was followed by an
11
etch back step to selectively remove the sacrificial top metal layer clearing the aperture geometry.
A 30 s etch with hydrofluoric acid diluted 1:10 in deionized water was then used to clear out the
residual HSQ in the antennas. Finally, the antenna structures were embedded in an UV curable ad-
hesive polymer (OrmoComp, Microresiste Technology GmbH) and stripped away from the silicon
wafers. The narrowest gap region lying at the bottom of the structure due to the metal diffusion
during the evaporation, was now flipped over to maximize the contact with the sample providing
direct accessibility to the plasmonic antenna hotspot. This method is fully scalable and shows ex-
cellent geometry control and planarity (see Supporting Information Fig. S1).
Cell culture, Atto647N-labeling and Cholesterol depletion of CHO cells CHO cells were
seeded on a coverslip containing planar nano-antennas with surface nanogaps and were allowed
to grow and spontaneously attach at 37o C in a controlled atmosphere with 5% of CO2 for nearly
48 hours. Lipid conjugates were separately prepared by labeling 1,2-Dipalmitoyl-sn-glycero-3-
phosphoethanolamine (DPPE) and Sphingomyelin (SM) with the organic dye Atto647N (from
Invitrogen) as described in Ref. 21 Prior to the fluorescence experiments, the lipid analogues were
incorporated in the cell membrane during a 3 mins incubation period at room temperature dis-
solved in the corresponding medium for CHO cells (Ham's F12 nutrient mixture). Stained cell
cultures were rinsed and washed to remove residual dye molecules before placing the sample cov-
erslip on the piezo-stage of an inverted microscope to carry out the measurements. For cholesterol
depletion experiments, the CHO cells were incubated in serum free buffer with 10 mM methyl-
β -cyclodextrin (MCD) for 30 mins at 37oC, and then the fluorescent labeling was carried out as
previously described. All fluorescence stainings were performed at a ∼300 nM concentration of
Atto647N and the measurements were completed within 30 mins after the incorporation of the
fluorescent analogs. From the number of detected fluorescence bursts (Fig. 2c) and the FCS am-
plitude, we estimate that the density of fluorescent lipids for the antenna experiments is on the
order of 20 to 80 probes per µ m2.
12
Experimental Setup The experiments were performed with a commercial MicroTime 200
setup equipped with an inverted confocal microscope (Olympus 60×, 1.2 NA water-immersion
objective) and a three-axis piezoelectric stage (PhysikInstrumente, Germany) allowing to select
individual nano-antennas. A linearly polarized 640 nm picosecond laser diode (Pico-Quant LDH-
D-C-640) in continuous wave mode was used to resonantly excite individual nano-antennas with
the laser linear polarization aligned with the main axis of the antenna dimer. The emitted fluo-
rescence signal was collected in epi-detection mode through a dichroic mirror and the signal was
finally split into two avalanche photodiodes (PicoQuant MPD-50CT). An emission filter and a
band pass 650-690 nm filter just before each detector was used to eliminate the scattered light by
the excitation laser. A 30 µ m pinhole in the detection arm yielded 0.5 fl confocal detection volume
at the sample plane. The fluorescence time traces were recorded on a fast time-correlated sin-
gle photon counting module in the time-tagged time-resolved mode (PicoQuant MPD-50CT). All
the fluorescence measurements were performed by illuminating the sample at an excitation power
density of ∼ 2-3 kW/cm2. The measurements were acquired for a typical run time of 50 s and the
correlation amplitudes were computed for ∼ 20 s windows with the commercial software package
SymPhoTime 64. Cells were cultured on different antenna samples, each sample containing dif-
ferent gap sizes.
Fluorescence Burst Analysis Single-molecule fluorescence time traces were acquired in the
Tagged Time-Resolved (TTTR) mode (recording each event at its arrival time) with 4 ps temporal
resolution. Fluorescence bursts analysis was then carried out with a likelihood-based algorithm
to test the null hypothesis (no burst, recording compatible with background noise) against the hy-
pothesis that a single-molecule burst arises as a consequence of a molecule crossing the excitation
area. Probabilities associated to false positive and missing event errors were both set to 10−3. 49
Fluorescence Correlation Spectroscopy The temporal fluctuations of the fluorescence inten-
sity F (t ) around the average value were analyzed to compute the temporal correlation G(τ ) =
13
⟨δ F (t ).δ F (t + τ )⟩/⟨F (t )⟩ , where δ F (t ) = F (t ) − ⟨F (t )⟩ is the fluctuation of the fluorescence
2
signal arising due to the molecules crossing the detection volume mediated by Brownian diffusion,
τ is the delay (lag) time, and ⟨ ⟩ indicates time averaging. The mobility of molecules shows strong
dependence on the local environment and thus in living systems the concept of ideal Brownian
diffusion may not always hold true. Considering possible anomalous diffusion in living cells, the
temporal correlation of the fluorescence intensity F can be written as: 18
where τdi f f (i) the average residence time of the it h diffusing modality, ρ (i) denotes the respective
amplitude contribution and α (i) being anomaly parameter of the same. 21 We find that the FCS
curves recorded with a nanoscopic illumination can only be fitted with a model assuming two
different diffusion modalities (i.e. ndiff = 2).
To define the probe areas used in the FCS diffusion laws (Fig. 4a-c), we use the product of the
gap size (measured by TEM) by the full width at half maximum for the intensity profile along the
direction perpendicular to the antenna main axis (computed by FDTD), following a calibration for
model lipid membranes. 44 Therefore, 10 nm and 35 nm gap sizes are associated respectively to
300 and 1250 nm2 probe areas.
Supporting Information
Supporting Information available: AFM image of antenna array, overlap between antenna's res-
onance and fluorescence spectra, fluorescence data for PE on different nano-antennas, overlay of
FCS curves from different nanoantennas, representative time trace of cholesterol depleted SM,
overlay of the confocal FCS data for PE and SM, FDTD simulations of intensity distributions,
FCS diffusion
laws with
confocal data,
fitting parameters
for FCS
curves.
14
Additional information
CM present address: Universitat de Vic, Universitat Central de Catalunya (UVic-UCC), C. de la
Laura 13, 08500 Vic, Spain
The authors declare no competing financial interests.
Acknowledgement
The authors thank Merche Rivas Jiménez, Felix Campelo and Erik Garbacik for technical support
and fruitful discussions. The research leading to these results has received funding from the Eu-
ropean Commission's Seventh Framework Programme (FP7-ICT-2011-7) under grant agreements
ERC StG 278242 (ExtendFRET) and 288263 (NanoVista). Financial support by the Spanish Min-
istry of Economy and Competitiveness ("Severo Ochoa" Programme for Centres of Excellence in
R&D (SEV-2015-0522) and FIS2014-56107-R grants) and Fundacion Privada Cellex is gratefully
acknowledged. RR is supported by the Erasmus Mundus Doctorate Program Europhotonics (Grant
159224-1-2009-1-FR-ERA MUNDUS-EMJD). PMW is supported by the ICFOstepstone Fellow-
ship, a COFUND Doctoral Program of the Marie Skłodowska-Curie Action from the European
Commission. CM acknowledges funding from the Spanish Ministry of Economy and Competi-
tiveness and the European Social Fund (ESF) through the Ramón y Cajal program 2015 (RYC-
2015-17896).
References
(1) Brown, D. A.; London, E. J. Biol. Chem., 2000, 275, 17221-17224.
(2) Lingwood, D.; Simons, K. Science, 2010, 327, 46-50.
(3) Kusumi, A.; Nakada, C.; Ritchie, K.; Murase, K.; Suzuki, K.; Murakoshi, H.; Kasai, R. S.;
Kondo, J.; Fujiwara, T.; Annu. Rev. Biophys. Biomol. Struct., 2005, 34, 351-378.
(4) Gowrishankar, K.; Ghosh, S.; Saha, S. C. R.; Mayor, S.; Rao, M. Cell 2012, 149, 1353-1367.
15
(5) Sezgin, E.; Levental, I.; Mayor, S.; Eggeling, C. Nat. Rev. Mol. Cell Biol., 2017, 18, 361-374.
(6) Mayor, S.; Rao, M. Traffic, 2004, 5, 231-240.
(7) Simons, K.; Gerl, M. J. Nat. Rev. Mol. Cell Biol., 2010, 11, 688-699.
(8) Hancock, J. F. Nat. Rev. Mol. Cell Biol., 2006, 7 456-62.
(9) Pike, L. J. J. Lipid Res. 2006, 47, 1597-1598.
(10) Lenne, P. F.; Wawrezinieck, L.; Conchonaud, F.; Wurtz, O.; Boned, A.; Guo, X. J.; Rigneault,
H.; He, H. T.; Marguet, D. EMBO J. 2006, 25, 3245-3256.
(11) Marguet, D.; Lenne, P. F.; Rigneault, H.; He, H. T. EMBO J. 2006, 25, 3446-3457.
(12) Wu, H.-M.; Lin, Y.-H.; Yen, T.-C.; Hsieh, C.-L. Sci. Rep. 2016, 6, 20542.
(13) van Zanten, T. S.; Cambi, A.; Koopman, M.; Joosten, B.; Figdor, C. G.; Garcia-Parajo, M. F.
Proc. Natl. Acad. Sci. USA., 2009, 106, 18557-18562.
(14) van Zanten, T. S.; Gómez, J.; Manzo, C.; Cambi, A.; Buceta, J.; Reigada, R.; Garcia-Parajo,
M. F. Proc. Natl. Acad. Sci. USA., 2010, 107, 15437-15442.
(15) Meder, D. M.; Moreno, J.; Verkade, P.; Vaz, W. L. C.; Simons, K. Proc. Natl. Acad. Sci.
USA., 2006, 103, 329-334.
(16) Dietrich, C.; Yang, B.; Fujiwara, T.; Kusumi, A.; Jacobson, K. Biophys. J., 2002, 82, 274-
284.
(17) Maiti, S.; Haupts, U.; Webb, W. W. Proc. Natl. Acad. Sci. USA., 1997, 94, 11753-11757.
(18) Bacia, K.; Kim, S. A.; Schwille, P. Nat. Methods, 2006, 3, 83-89.
(19) Wawrezinieck, L.; Rigneault, H.; Marguet, D.; Lenne, P.-F. Biophys. J., 2005, 89, 4029-4042.
(20) Kastrup, L.; Blom, H.; Eggeling, C.; Hell, S.W. Phys. Rev. Lett., 2005, 94(17), 1-4.
16
(21) Eggeling, C.; Ringemann, C.; Medda, R.; Schwarzmann, G.; Sandhoff, K.; Polyakova, S.;
Belov, V. N.; Hein, B.; von Middendorff, C.; Schönle, A.; Hell, S. W. Nature, 2009, 457,
1159-1162.
(22) Mueller, V.; Ringemann, C.; Honigmann, A. ; Schwarzmann, G.; Medda, R.; Leutenegger,
M.; Polyakova, S.; Belov, V. N.; Hell, S. W.; Eggeling, C. Biophys. J., 2011, 101, 1651-1660.
(23) Honigmann, A.; Mueller, V.; Ta, H.; Schoenle, A.; Sezgin, E.; Hell, S. W.; Eggeling, C. Nat.
Commun. 2014, 5, 5412.
(24) Vicidomini, G.; Ta, H.; Honigmann, A. ; Mueller, V.; Clausen, M. P.; Waithe, D.; Galiani, S.;
Sezgin, E.; Diaspro, A.; Hell, S. W.; Eggeling, C. Nano Lett., 2015, 15, 5912-5918.
(25) Punj, D.; Ghenuche, P.; Moparthi, S. B.; de Torres, J.; Grigoriev, V.; Rigneault, H.; Wenger,
J. WIREs Nanomed. Nanobiotechnol., 2014, 6, 268-282.
(26) Edel, J. B.; Wu, M.; Baird, B.; Craighead, H. G. Biophys. J. 2005, 88, L43-L45.
(27) Moran-Mirabal, J. M.; Torres, A. J.; Samiee, K. T.; Baird, B. A.; Craighead, H. G. Nanotech-
nology 2007, 18, 195101.
(28) Wenger, J.; Conchonaud, F.; Dintinger, J. Wawrezinieck, L.; Ebbesen, T. W.; Rigneault, H.;
Marguet, D.; Lenne, P.-F. Biophys. J., 2007, 92, 913-919.
(29) Kelly, C. V.; Baird, B. A.; Craighead, H. G. Biophys. J. 2011, 100, 34-36.
(30) Kelly, C. V.; Wakefield, D. L.; Holowka, D. A.; Craighead, H. G.; Baird, B. A. ACS Nano
2014, 8, 7392-7404.
(31) Richards, C. I.; Luong, K.; Srinivasan, R.; Turner, S. W.; Dougherty, D. A.; Korlach, J.;
Lester, H. A. Nano Lett. 2012, 12, 3690-3694.
(32) Lohmüller, T.; Iversen, L.; Schmidt, M.; Rhodes, C.; Tu, H. L.; Lin, W. C.; Groves, J. T.
Nano Lett. 2012, 12, 1717-1721.
17
(33) Mivelle, M.;, Van Zanten, T. S; Neumann, L.; van Hulst, N.; García-Parajo, M. F. Nano Lett.,
2012, 12 (11), 5972-5978.
(34) Flauraud, V.; van Zanten, T.S.; Mivelle, M.; Manzo, C.; García-Parajo, M.F.; Brugger, J.
Nano Lett. 2015, 15, 4176-4182.
(35) Pradhan, B.; Khatua, S.; Gupta, A.; Aartsma, T.; Canters, G.; Orrit, M. J. Phys. Chem. C
2016, 120, 25996-26003.
(36) García-Parajo, M. F. Nat. Photonics, 2008, 2, 201-203.
(37) Manzo, C.; Van Zanten, T. S; García-Parajo, M. F. Biophys. J., 2011, 1000, 08-10.
(38) Novotny, L.; van Hulst, N. Nat. Photonics 2011, 5, 83-90.
(39) Kinkhabwala, A.; Yu, Z. F.; Fan, S. H.; Avlasevich, Y.; Mullen, K.; Moerner, W. E. Nat.
Photonics 2009, 3, 654-657.
(40) Acuna, G. P.; Möller, F. M.; Holzmeister, P.; Beater, S.; Lalkens, B.; Tinnefeld, P. Science
2012, 338, 506-510.
(41) Punj, D.; Mivelle, M.; Moparthi, S.B.; van Zanten, T.S.; Rigneault, H.; van Hulst, N.F.;
García-Parajo, M.F.; Wenger, J. Nat. Nanotechnol. 2013, 8, 512-516.
(42) Puchkova, A.; Vietz, C.; Pibiri, E.; Wünsch, B.; Sanz Paz, M.; Acuna, G. P.; Tinnefeld, P.
Nano Lett. 2015, 15, 8354-8359.
(43) Flauraud, V.; Regmi, R.; Winkler, P. M.; Alexander, D. T. L.; Rigneault, H.; van Hulst, N. F.;
García-Parajo, M. F.; Wenger, J.; Brugger, J. Nano Lett. 2017, 17, 1703-1710.
(44) Winkler, P. M.; Regmi, R.; Flauraud, V.; Brugger, J.; Rigneault, H.; Wenger, J.; García-
Parajo, M. F. ACS Nano 2017, 11, 7241-7250.
(45) Langguth, L.; Koenderink, A. F. Opt. Express 2014, 22, 15397-15409.
18
(46) Nicolau, D. V.; Burrage, K.; Parton, R. G.; Hancock, J. F. Mol. Cell. Biol. 2006, 26, 313-323.
(47) Spillane, K. M.; Ortega-Arroyo, J.; de Wit, G.; Eggeling, C.; Ewers, H.; Wallace, M. I.;
Kukura, P. Nano Lett. 2014, 14, 5390-5397.
(48) Spindler, S.; Ehrig, J.; König, K.; Nowak, T.; Piliarik, M.; Stein, H. E.; Taylor, R. W.;
Garanger, E.; Lecommandoux, S.; Alves, I. D.; Sandoghdar, V. J. Phys. Appl. Phys. 2016, 49,
274002.
(49) Zhang, K.; Yang, H. J. Phys. Chem. B, 2005, 109, 21930-21937.
19
Figure 1: Planar gold nano-antenna arrays for probing single-molecule dynamics in the
plasma membrane of living cells. (a) CHO cells are seeded onto a microscopic coverslip con-
taining multiple planar nano-antennas with 10 and 35 nm gap sizes. The inset shows the cross
section of the antenna-in-box stripped and embedded into a polymer, bringing the region of max-
imum electromagnetic field intensity onto the surface in direct contact with the plasma membrane
of living cells. (b) From left to right: macro-photograph of a coverslip with a stripped Au film with
large-scale planar antenna arrays; dark field optical micrograph of a small portion of the antenna
arrays showing here 625 antennas with 10 nm nominal gap size; transmission electron microscope
(TEM) images of antennas with 10 and 35 nm gap size. (c) Confocal image of CHO cells showing
the morphology after incorporating the fluorescent SM lipid analog labeled with Atto647N.
20
Figure 2: Single-molecule fluorescence time traces in living CHO cells. (a,b) Fluorescence time
traces for phosphoethanolamine (PE, left) and sphingomyelin (SM, right) labeled with Atto647N
recorded with confocal (a) and with a 10 nm gap planar nano-antenna (b). The binning time is
0.1 ms for all traces. The diffraction-limited spot in the confocal configuration cannot resolve the
nanoscopic and heterogeneous membrane organization, thus results in indistinguishable fluores-
cence time traces for both PE and SM. However, the highly confined nano-antenna hotspot reveals
clear differences in the diffusion dynamics of PE and SM. (c) The fluorescence time traces are
analyzed to produce scatter plots showing the distribution of fluorescence burst intensity versus
burst duration. Single-molecule events in sub-ms time scales are observed with nano-antennas
(color dots) as the confined electromagnetic hotspots allow to probe the dynamics occurring be-
yond the diffraction limit. Single molecule events obtained with confocal illumination are shown
for comparison (black dots).
21
Figure 3: Nano-antenna FCS on living cell membranes. (a,b) Normalized fluorescence corre-
lation curves for Atto647N labeled PE (a) and SM (b) lipid analogs probed with nano-antennas
of varying gap size. The color lines are experimental data and the black curves are numerical
fits. Each FCS trace is a representative example taken on an individual nano-antenna. FCS curves
recorded on different nano-antennas and different cells are shown in Fig. S4. The diffraction-
limited confocal measurements are shown in gray for direct comparison. (c) Comparison of FCS
curves for PE and SM for a 10 nm gap antenna. Unlike the confocal reference, the nano-antenna re-
veals clear differences between the dynamics of PE and SM at the nanoscale. (d) After cholesterol
depletion, the SM diffusion dynamics are significantly faster and resemble the PE case.
22
Figure 4: Characteristic diffusion dynamics of membrane lipids probed with ultra-confined
nano-antenna hotspots. The diffusion time measured by FCS (for 60 different nano-antennas) is
plotted as a function of the probe area for PE (a), SM (b), and SM after MCD treatment(c). The
solid lines are linear fits through the median values. In the case of free diffusion, the origin (0, 0)
is aligned with the expected line, while a positive intercept at the y-axis denotes hindered diffusion
due to nanodomains. (d) The diffusion coefficients computed from the slopes in a-c are compared
with confocal results. (e) y-axis intercept deduced from the linear fits in a-c. PE and MCD-treated
SM show near-zero y-intercept consistent with free diffusion, while the significant y-intercept for
SM indicates that the diffusion is constrained by nanodomains. (f) Normalized spread in diffusion
time (width of upper and lower quartiles / median) in each case. The large dispersion observed is
SM is another indication that sphingolipids are preferentially recruited into transient nanoscopic
domains.
23
|
1004.3837 | 1 | 1004 | 2010-04-22T05:19:55 | Numerical study on the emergence of anisotropy in artificial flocks: A BOIDS modeling and simulations of empirical findings | [
"physics.bio-ph",
"nlin.AO"
] | In real flocks, it was revealed that the angular density of nearest neighbors shows a strong {\it anisotropic structure} of individuals by very recent extensive field studies by Ballerini et al [{\it Proceedings of the National Academy of Sciences USA} {\bf 105}, pp.1232-1237 (2008)]. In this paper, we show that this empirical evidence in real flocks, namely, the structure of anisotropy also emerges in an artificial flock simulation based on the {\it BOIDS} by Reynolds [{\it Computer Graphics} {\bf 21}, pp.25-34 (1987)]. We numerically find that appropriate combinations of the weights for just only three essential factors of the BOIDS, namely, `Cohesion', `Alignment' and `Separation' lead to a strong anisotropy in the flock. This result seems to be highly counter-intuitive and also provides a justification of the hypothesis that the anisotropy emerges as a result of self-organization of interacting intelligent agents (birds for instance). To quantify the anisotropy, we evaluate a useful statistics (a kind of {\it order parameters} in statistical physics), that is to say, the so-called $\gamma$-value defined as an inner product between the vector in the direction of the lowest angular density of flocks and the vector in the direction of the moving of the flock. Our results concerning the emergence of the anisotropy through the $\gamma$-value might enable us to judge whether an arbitrary flock simulation seems to be {\it realistic} or not. | physics.bio-ph | physics |
Numerical study on the emergence of anisotropy in artificial flocks:
A BOIDS modeling and simulations of empirical findings
Motohiro Makiguchi and Jun-ichi Inoue
Complex Systems Engineering, Graduate School of Information Science and Technology
Hokkaido University, N14-W9, Kita-Ku, Sapporo 060-0814, Japan
November 29, 2009
Abstract
In real flocks, it was revealed that the angular density of nearest neighbors shows a strong anisotropic structure of
individuals by very recent extensive field studies by Ballerini et al [Proceedings of the National Academy of Sciences
USA 105, pp.1232-1237 (2008)]. In this paper, we show that this empirical evidence in real flocks, namely, the
structure of anisotropy also emerges in an artificial flock simulation based on the BOIDS by Reynolds [Computer
Graphics 21, pp.25-34 (1987)]. We numerically find that appropriate combinations of the weights for just only
three essential factors of the BOIDS, namely, 'Cohesion', 'Alignment' and 'Separation' lead to a strong anisotropy
in the flock. This result seems to be highly counter-intuitive and also provides a justification of the hypothesis
that the anisotropy emerges as a result of self-organization of interacting intelligent agents (birds for instance). To
quantify the anisotropy, we evaluate a useful statistics (a kind of order parameters in statistical physics), that is
to say, the so-called γ-value defined as an inner product between the vector in the direction of the lowest angular
density of flocks and the vector in the direction of the moving of the flock. Our results concerning the emergence
of the anisotropy through the γ-value might enable us to judge whether an arbitrary flock simulation seems to be
realistic or not.
keywords: Self-organization, Anisotropy, BOIDS, Swarm Intelligence Simulation, Collective behaviour
1 Introduction
Collective behaviour of interacting intelligent agents such as birds, insects or fishes shows highly non-trivial properties
and sometimes it seems to be quite counter-intuitive [1]. As well-known, many-body systems having a lot of non-
intelligent elements, for instance, spins (tiny magnets in atomic scale length), particles, random-walkers etc. also show
a collective behaviour like a critical phenomenon of order-disorder phase transitions with 'spontaneous symmetry
breaking' in spatial structures of the system. Up to now, a huge number of numerical studies in order to figure it out
have been done by theoretical physicists and mathematicians [2]. They attempted to describe these phenomena by
using some probabilistic models and revealed the 'universality class' of the critical phenomena by solving the problem
with the assistance of computer simulations. Of course, the validity of the studies should be checked by comparing
the numerical results with the experimental findings. If their results disagree with the empirical data, the models they
used should be thrown away or should be modified appropriately.
On the other hand, for the mathematical modeling of many-body systems having interacting intelligent agents
(animals), we also use some probabilistic models, however, it is very difficult for us to evaluate the modeling and also
very hard to judge whether it looks like realistic or not due to a lack of enough empirical data to be compared.
One of the key factors for such non-trivial collective behaviour of both non-intelligent and intelligent agents is
obviously a 'competition' between several different (and for most of the cases, these are incompatible) effects. For
instance, the Ising model as an example of collective behaviour of non-intelingent agents exhibits an order-disorder
phase transition [2] by competition between the ferromagnetic interactions between Ising spins ('energy minimization')
and thermal fluctuation ('entropy maximization') by controlling the temperature of the system. On the other hand,
as a simplest and effective algorithm in computer simulations for flocks of intellingent agents, say, animals such as
starlings, the so-called BOIDS founded by Reynolds [3, 4] has been widely used not only in the field of computer
graphics but also in various other research fields including ethology, physics, control theory, economics, and so on.
1
The BOIDS simulates the collective behaviour of animal flocks by taking into account only a few simple rules for each
interacting intelligent agent.
However, there are few studies to compare the results of the BOIDS simulations with the empirical data. Therefore,
the following essential and interesting queries still have been left unsolved;
• What is a criterion to determine to what extent the flocks seem to be realistic?
• Is there any quantity (statistics) to measure the quality of the artificial flocks?
From the view point of 'engineering', the above queries are (in some sense) not essential because their main goal is
to construct a useful algorithm based on the collective behaviour of agents. However, from the natural science view
points, the difference between empirical evidence and the result of the simulation is the most important issue and the
consistency is a guide to judge the validity of the computer modeling and simulation.
Recently, Ballerini at al [5] succeeded in obtaining the data for such collective animal behaviour, namely, empirical
data of starling flocks containing up to a few thousands members. They also pointed out that the angular density of
the nearest neighbors in the flocks is not uniform but apparently biased (it is weaken) along the direction of the flock's
motion.
With their empirical findings in mind, in this paper, we examine the possibility of the BOIDS simulations to
reproduce this anisotropy and we also investigate numerically the condition on which the anisotropy emerges.
This paper is organized as follows. In the next section, we explain the empirical findings by Ballerini et al [5, 6]
and introduce a key concept anisotropy and a relevant quantity γ-value. In section 3, the BOIDS modeling and setting
of essential parameters in our simulations are explicitly explained. The results are reported in section 4. The last
section provides concluding remarks.
2 Empirical findings by Ballerini et al
In this section, we briefly review the measurement of the realistic flocks and the evaluation of the empirical data by
Ballerini et al [5]. They measured each bird's position in the flocks of starling (Sturnus vulgaris) in three dimension.
To get such 3D data, they used 'Stereo Matching' which reconstructs 3D-object from a set of stereo photographs.
2.1 Anisotropy
From these data, they calculated the angular density of the nearest neighbours in the flock. They measured the
angles (φ, α), where φ stands for the 'latitude' of the nearest neighbour for each bird measured from the direction of
the motion of the flock, whereas α denotes 'longitude' which specifies the position of the nearest neighbour for each
bird around the direction of flock's motion, for all individuals in the flock and made the 2D-map of angular density
distribution using the so-called 'Mollweide projection'. Their figure clealy shows that the density is not uniform but
obviously biased. For instance, we find from the figure that the dinsity around φ (cid:39) 0 and α (cid:39) 0o,±180o are extremely
low in comparison with the density in the other directions. The property of the biased distribution due to the absence
of the birds along the direction of the flock's motion is referred to as anisotropy [5]. The main goal of this paper is
to reveal numerically that the artificial flock by the BOIDS exhibits the anisotropy as the realistic flock shows [5]. To
quantify the degree of the anisotropy, we use a useful statistics (a kind of 'order parameters' in the research field of
statistical mechanics) introduced in the following subsections.
2.2 The γ-value: An order parameter to detect 'spatial symmetry breaking'
Ballerini et al also introduced a useful indicator, what we call γ-value. The γ-value is calculated according to the
following recipe. Let u(n)
be an unit vector pointing in the direction of the nth -nearest neighbour of the bird i and
let us define the projection matrix M (n) in terms of the u(n)
as follows.
i
ui
ui
ui
(n)
x
(n)
y
(n)
z
, (M (n))αβ =
i
N(cid:88)
i=1
1
N
u(n)
i =
(u(n)
i
)α(u(n)
i
)β (α, β = x, y, z)
where N is the number of birds in the flock. Then, the γ-value is given by
γ = (cid:104)(W (n) · V )
2(cid:105)
(1)
2
where W (n) denotes the normalized eigenvector corresponding to the smallest eigenvalue of the projection matrix
M (n). From the definition, the W (n) coincides with the direction of the lowest density in the flock. The vector V
appearing in the equation (1) means the unit vector of flock's motion. The bracket (cid:104)···(cid:105) means the average over the
ensembles of the flocks. The γ-value for the uniform distribution of the position (ϕ, θ) for a given vector V , namely,
the γ for ρ(ϕ, θ) = (4π)−1 is easily calculated as
(cid:90) 2π
0
(cid:90) π
−π
γisotropy =
dϕ
dθρ(ϕ, θ) sin θ cos2 θ =
1
3
where we used (W (n) · V )
= cos2 θ. Therefore, the distribution of the nth-nearest neighbours has anisotropic structure
when the γ-value is larger than 1/3, namely the condition for the emergence of the anisotropy is explicitly written by
2
γ > γisotropy =
1
3
.
(2)
By measuring this γ-value for artificial flock simulations, one can show that the anisotropy also emerges in computer
simulations. To put it into other words, the system of flocks is spatially 'symmetric' for γ = γisotropy, whereas the
symmetry is 'spontaneously' broken for γ > γisotropy. This 'spontaneous symmetry breaking' is nothing but the
emergence of anisotropy.
In the following sections, we carry out the BOIDS simulations and evaluate the anisotropy by the γ-value. Then,
we find that the 'spontaneous symmetry breaking' mentioned above actually takes place by controlling the essential
parameters appearing in the BOIDS.
3 The BOIDS modeling and simulations
To make flock simulations in computer, we use the so-called BOIDS which was originally designed by Reynolds [3].
The BOIDS is one of the well-known mathematical (probabilistic) models in the research fields of CG and animation.
Actually, the BOIDS can simulate very complicated animal flocks or schools although it consists of just only three
simple interactions for each agent in the aggregation:
(c) Cohesion: Making a vector of each agent's position toward the average position of local flock mates.
(a) Alignment: Making a vector of each agent's position towards the average heading of local flock mates and
keeping the velocity of each agent the average value of flock mates.
(s) Separation: Controlling the vector of each agent to avoid the collision with the other local flock mates.
It is important for us to bear in mind that 'local flock mates' mentioned above denotes the neighbours within the range
of views for each agent. Each agent decides her (or his) next migration by compounding these three interactions.
3.1 On the setting of essential parameters in BOIDS simulations
In our BOIDS simulations, each agent is defined as a mass point and specified by a set of 3D-coordinate (x, y, z), an
unit vector of motion, and the speed. We define each agent's view as a sphere with a radius R without any blind
corner. We also define the 'separation sphere' with a radius R0 and the distance between the nearest neighbours is
specified by the length D1 (see Figure 1). The interaction of 'Separation' is switched on if and only if the D1 is smaller
than the R0. Some other essential parameters appearing in our BOIDS simulations and the setup are also explicitly
given as follows.
there is no wall and no ground surface.
• Field of simulations: Three dimensional open space without any gravity or any air resistance. Moreover,
• The number of agents in the flock: The system size of simulations is N = 100
• Initial condition on the speed of each agent: 1 < speed < 8.
• Initial condition on the location of each agent: All agents are distributed in a sphere with radius L = 180.
• The shape and the range of each agent's view: A sphere with radius R = 200.
• The shape and the range of separation: A sphere with radius R0 = 20.
For the above setting of the parameters, we shall implement two types of programming codes. One is a programing
code for a single simulation ('SS' for short), and another code is for multiple simulations ('MS' for short). The SS
runs in the GUI (graphical user interface) and it shows us a shape of the flock in real time, whereas the MS enables
us to carry out a number of simulations with different initial conditions.
3
Figure 1: Range of View (R) and Separation(R0).
3.2 Typical four aggregations
By controling the three essential interactions, namely, 'Cohesion', 'Alignment' and 'Separation' mentioned above,
we obtain four different aggregations having different collective behaviours. Each behavour of the aggregations is
monitored (observed) by the SS. To specify the aggregation process of the flocks, we define the update rule of the
vector of movement for each agent by
v(t + 1) =
P0V Cohesion(t) + P1V Alignment(t) + P2V Separation(t) + P3v(t)
P0V Cohesion(t) + P1V Alignment(t) + P2V Separation(t) + P3v(t)
(3)
where t means the time step of the update and A denotes the L1-norm of a vector A. V Cohesion is a vector pointing
to the center of mass from each agent's position. V Alignment denotes a vector to be obtained by averaging over the
velocities of all agents. V Separation means a vector pointing to the direction of the movement of each agent to be
separated from her (or his) nearest neighbouring mate. Therefore, the aggregation of the flock is completely specified
by the weights of the above vectors, namely, P ≡ (P0, P1, P2, P3). Among all possible combinations of these weights
P , we shall pick up typical four cases. Each property and the shape of each aggregation are explained as follows.
• Case 1 (Crowded Aggregation): The aggregation obtained by controlling the interaction of 'Cohesion' much
stronger than the others, namely, P = (1, 0, 0, 1).
• Case 2 (Spread Aggregation): The aggregation obtained by controlling the interaction of 'Separation' much
stronger than the others, namely, P = (0, 0, 1, 1).
• Case 3 (Synchronized Aggregation): The aggregation obtained by controlling the interaction of 'Alignment'
much stronger than the others, namely, P = (0, 1, 1, 1).
• Case 4 (Flock Aggregation): This aggregation obtained by adjusting every interactions appropriately, namely,
P = (1, 5, 1.5.1).
Using the MS, we simulate each aggregation for 200 times for different initial conditions.
In each simulation, we
measure the angular distribution and then the value of γ is calculated. The number of crash is also updated when the
coordinate of each agent is identical to (is shared with) the other agents. We start each measurement from the time
point at which the total amount of the change in every agent's speed is close to 0 through the 80 turns of the update.
In the next section, we explain the details of the result.
4 Results
In association with the each aggregation, we evaluate the γ-value with standard deviation and the average number of
crashes for 200 independent runs of the BOIDS simulations. In following, we summarize the results.
• Case 1 (Crowded Aggregation): The typical behaviour of the flock and the angular distribution are shown in
the upper left panel in Figure 2 and Figure 3, respectively. The γ-value is 0.333 with standard deviation 0.083
and the average number of crashes is 2091.74.
4
(cid:7536)(cid:7536)(cid:7502)(cid:7522)(cid:7503)• Case 2 (Spread Aggregation): The typical behaviour of the flock and the angular distribution are shown in
the upper right panel in Figure 2 and Figure 3, respectively. The γ-value is 0.315 with standard deviation 0.258
and the average number of crashes is 0.
• Case 3 (Synchronized Aggregation): The typical behaviour of the flock and the angular distribution are
shown in the lower left panel in Figure 2 and Figure 3, respectively. The γ-value is 0.300 with standard deviation
0.281 and the average number of crashes is 0.
• Case 4 (Flock Aggregation): The typical behaviour of the flock and the angular distribution are shown in the
lower right panel in Figure 2 and Figure 3, respectively. The γ-value is 0.744 with standard deviation 0.124 and
the average number of crashes is 0.
Figure 2: From the upper left to the lower right, snapshots of the typical behaviour for 'Crowded Aggregation', 'Spread Aggregation',
'Synchronized Aggregation' and 'Flock Aggregation' (the projections to the xy- and xz-planes) are shown.
The aggregation of Case 4 ('Flock Aggregation') has the highest γ-value among the four cases and its angular density
clearly shows a lack of nearest neighbours along the direction of flock's motion leading to an anisotropy. For all
aggregations except for the Case 4, the γ-values are lower than γisotropy = 1/3, namely, they have no anisotropy. From
these results, we find that BOIDS computer simulations having appropriate weights P shows anisotropy structures as
real flocks exhibit. It is also revealed that one can evaluate to what extent an arbitrary flock simulation is close to
real flocks through the γ-value.
From the results obtained here, we might have another question, namely, it is important for us to answer the
question such as whether the aggregation having a higher γ-value than the Case 4 seems to be more realistic than
the Case 4 or not. To answer the question, we carry out the simulations of the flock aggregation which has a higher
γ-value (Case 5) than the Case 4. The results are summarized as follows.
• Case 5 (Crowded Aggregation): The angular distribution is shown in Figure 4. The γ-value is 0.931 with
standard deviation 0.02 and the average number of crashes is 570.6.
Obviously, the above aggregation has the highest γ-value leading to the strongest anisotropy among the five cases.
However, it is hard for us to say that it is an optimal flock because the number of cashes is also the highest (570 times)
and to make matter worse, the number itself is apparently outstanding. This result tells us that an aggregation having
much stronger anisotropic structures is not always a better flock.
The above result is reasonably accepted because the γ-value is calculated from the angular distribution of nearest
neighbours without any concept of the distance between agents. Therefore, the flock having a dense network might have
highly risks of crashes more than the sparse network. For this reason, in order to judge whether a given aggregation
has a better flock behaviour or not, we should use the other criteria which take into account the distance between
nearest neighbours.
5
xy_planexz_planexy_planexz_planexy_planexz_planexy_planexz_planeFigure 3: From the upper left to the lower right, angular density of the Case 1 ( 'Crowded Aggregation'), the Case 2 ('Spread Aggregation'),
the Case 3 ('Synchronized Aggregation') and the Case 4 ('Flock Aggregation') are shown. The spatial symmetry in the Case 4 is apparently
broken (the anisotropy emerges).
Figure 4: The angular density for the aggregation (Case 5) having a higher γ-value than the Case 4.
Inspired by the empirical data analysis by Ballerini et al [5], we finally calculate the γ-value as a function of the
order of the neighbour. The result is shown in Figure 5. In this figure, n denotes the order of the neighbour, for
instance, n = 1 or n = 2 means the nearest neighbour, the next nearest neighbour, respectively. The figure shows
a similar behaviour to the corresponding plot in the reference [5], that is, the γ-values monotonically decrease as n
increases and they converge to γisotropy = 1/3 beyond n (cid:39) 6. This result might be a justification to conclude that our
BOIDS simulations having appropriate weight vectors P actually simulate a realistic flock.
5 Concluding remarks
In this paper, we showed that the anisotropy observed in the empirical data analysis [5] also emerges in our BOIDS
simulations having appropriate weight vectors P . From the γ-value we calculated, one can judge wheter an optional
aggregation behaves like a real flock or not. The system of flocks is spatially 'symmetric' for γ = γisotropy = 1/3,
whereas the symmetry is 'spontaneously' broken for γ > γisotropy. We found from the behaviour of 'order parameter'
γ that this 'spontaneous symmetry breaking' is nothing but the emergence of anisotropy.
As well-known, there are some conjectures on the origin of the emergence of anisotropy. For instance, the effect of
bird's vision is one of the dominant hypotheses. In fact, real starlings have lateral visual axes and each of the starlings
has a blind rear sector [8]. If all individuals in the flock move to avoid their nearest neighbours which are hidden
in their blind sectors, the effect of the blind sector is more likely to be a factor to emerge the anisotropy of nearest
neighbours in the front-rear directions. However, our result proved that this hypothesis is NOT ALWAYS correct
because agents in our simulation have no blind sector of their views. Nevertheless, we found that our flock aggregation
6
+90(cid:15078)-90(cid:15078)+180(cid:15078)-180(cid:15078)+90(cid:15078)-90(cid:15078)+180(cid:15078)-180(cid:15078)+90(cid:15078)-90(cid:15078)+180(cid:15078)-180(cid:15078)+90(cid:15078)-90(cid:15078)+180(cid:15078)-180(cid:15078)+90(cid:15078)-90(cid:15078)+180(cid:15078)-180(cid:15078)Figure 5: The γ-value as a function of the neighbour n in the real flock.
has an anisotropy of the nearest neighbours. The result means that the agent's blind as an effect of vision is not
necessarily required to produce the anisotropy and much more essential factor for the anisotropy is the best possible
combinations of three essential interactions in the BOIDS.
We hope that these results might help us to consider the relevant link between BOIDS simulations and empirical
evidence from real world.
Acknowledgement
We were financially supported by Grant-in-Aid Scientific Research on Priority Areas 'Deepening and Expansion of
Statistical Mechanical Informatics (DEX-SMI)' of the MEXT No. 18079001. One of the authors (JI) was financially
supported by INSA (Indian National Science Academy) - JSPS (Japan Society of Promotion of Science) Bilateral
Exchange Programme. He also thanks Saha Institute of Nuclear Physics for their warm hospitality during his stay in
India.
References
[1] Iwao Bialynicki-Birula and Iwona Bialynicka-Birula, Modeling Reality: How computers mirror life, Oxford Uni-
versity Press (2004).
[2] D.P. Landau and K. Binder, A Guide to Monte Carlo Simulations in Statistical Physics, Cambridge University
Press (2000).
[3] C.W. Reynolds, Flocks, Herds, and Schools: A Distributed Behavioral Model, Computer Graphics 21, pp.25-34
(1987).
[4] http://www.red3d.com/cwr/boids/
[5] M. Ballerini, N. Cabibbo, R. Candelier, A. Cavagna, E. Cisbani, I. Giardina, V. Lecomte, A. Orlandi, G. Parisi, A.
Procaccini, M. Viale and V. Zdravkovic, Interaction Rulling Animal Collective Behaviour Depends on Topological
raher than Metric Distance, Evidence from a Field Study, Proceedings of the National Academy of Sciences USA
105, pp.1232-1237 (2008).
[6] A. Cavagna, I. Giardina, A. Orlandi, G. Parisi, A. Procaccini, M. Viale and V. Zdravkovic, The STARFLAG
handbook on collective animal behaviour: Part II, empirical methods, Animal Behaviour 76, Issue 1, pp237-248
(2008).
[7] J.M. Cullen, E. Shaw and H.B. Baldwin, Methods for measuring the three dimensional structure of fish schools,
Animal Behaviour 13, pp. 534-543 (1965).
[8] G.R. Martin, The eye of a passeriform bird, the European starling (Sturnus vulgaris): eye movement amplitude,
visual fields and schematic optics, J. Comp. Physiol. A 159, pp. 545-557 (1986).
7
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 0 2 4 6 8 10 12 14 16(cid:97)nBOIDS simulation(cid:97)isotropy = 1/3 |
1701.02565 | 1 | 1701 | 2017-01-05T22:09:48 | The Energetic Cost of Building a Virus | [
"physics.bio-ph",
"q-bio.PE",
"q-bio.SC"
] | Viruses are incapable of autonomous energy production. Although many experimental studies make it clear that viruses are parasitic entities that hijack the host's molecular resources, a detailed estimate for the energetic cost of viral synthesis is largely lacking. To quantify the energetic cost of viruses to their hosts, we enumerated the costs associated with two very distinct but representative DNA and RNA viruses, namely T4 and influenza. We found that for these viruses, translation of viral proteins is the most energetically expensive process. Interestingly, the cost of building a T4 phage and a single influenza virus are nearly the same. Due to influenza's higher burst size, however, the overall cost of a T4 phage infection is only 2-3% of the cost of an influenza infection. The costs of these infections relative to their host's estimated energy budget during the infection reveal that a T4 infection consumes about a third of its host's energy budget, where as an influenza infection consumes only 1%. Building on our estimates for T4, we show how the energetic costs of double-stranded DNA viruses scale with virus size, revealing that the dominant cost of building a virus can switch from translation to genome replication above a critical virus size. Lastly, using our predictions for the energetic cost of viruses, we provide estimates for the strengths of selection and genetic drift acting on newly incorporated genetic elements in viral genomes, under conditions of energy limitation. | physics.bio-ph | physics | The Energetic Cost of Building a Virus
Gita Mahmoudabadi1, Ron Milo2, Rob Phillips1,3
1 Department of Bioengineering, California Institute of Technology, Pasadena, CA
91125, USA. 2 Department of Plant and Environmental Sciences, Weizmann Institute of
Science, Rehovot 7610001, Israel. 3 Department of Applied Physics, California Institute
of Technology, Pasadena, CA 91125, USA.
Abstract
Viruses are incapable of autonomous energy production. Although many experimental studies
make it clear that viruses are parasitic entities that hijack the host's molecular resources, a
detailed estimate for the energetic cost of viral synthesis is largely lacking. To quantify the
energetic cost of viruses to their hosts, we enumerated the costs associated with two very
distinct but representative DNA and RNA viruses, namely, T4 and influenza. We found that for
these viruses, translation of viral proteins is the most energetically expensive process.
Interestingly, the cost of building a T4 phage and a single influenza virus are nearly the same.
Due to influenza's higher burst size, however, the overall cost of a T4 phage infection is only 2 -
3% of the cost of an influenza infection. The costs of these infections relative to their host's
estimated energy budget during the infection reveal that a T4 infection consumes about a third
of its host's energy budget, whereas an influenza infection consumes only 1%. Building on our
estimates for T4, we show how the energetic costs of double-stranded DNA viruses scale with
virus size, revealing that the dominant cost of building a virus can switch from translation to
genome replication above a critical virus size. Lastly, using our predictions for the energetic cost
of viruses, we provide estimates for the strengths of selection and genetic drift acting on newly
incorporated genetic elements in viral genomes, under conditions of energy limitation.
Significance Statement
Viruses rely entirely on their host as an energy source. Despite numerous experimental studies
that demonstrate the capability of viruses to rewire and undermine their host's metabolism, we
still largely lack a quantitative understanding of an infection's energetics. And yet, the energetics
of a viral infection is at the center of broader evolutionary and physical questions in virology. By
enumerating the energetic costs of different viral processes, we open the door to quantitative
predictions about viral evolution. For example, we predict that for the majority of viruses,
translation will serve as the dominant cost of building a virus, and that selection, rather than
drift, will govern the fate of new genetic elements within viral genomes.
Key Words
viral energetics, viral evolution, T4, influenza, cellular energetics
1
Introduction
Viruses are biological 'entities' at the boundary of life. Without cells to infect, viruses as we know
them would cease to function, as they rely on their hosts to replicate. Though the extent of this
reliance varies for different viruses, all viruses consume from the host's energy budget in
creating the next generation of viruses. There are many examples of viruses that actively
subvert the host transcriptional and translational processes in favor of their own replication (1).
This viral takeover of the host metabolism manifests itself in a variety of forms such as in the
degradation of the host's genome or the inhibition of the host's mRNA translation (1). These
examples suggest that a viral infection requires a considerable amount of the host's energetic
supply. In support of this view are experiments on T4 (2), T7 (3), Pseudoalteromonas phage (4),
and Paramecium bursaria chlorella virus-1 or PBCV-1 (5), demonstrating the viral burst size to
correlate positively with the host growth rate. In the case of PBCV-1, the burst size is reduced
by 50% when its photosynthetic host, a freshwater algae, is grown in the dark (5). Similarly,
slow growing E. coli with a doubling time of 21 hours affords a T4 burst size of just one phage
(6), as opposed to a burst size of 100-200 phages during optimal growth conditions.
There are many other experimental studies (discussed in the SI section I) (7-11) that
demonstrate viruses to be capable of rewiring the host metabolism. These fascinating
observations led us to ask the following questions: what is the energetic cost of a viral infection,
and what is the energetic burden of a viral infection on the host cell? To our knowledge, the first
attempt to address these problems is provided through a kinetic model of the growth of Qss
phage, which demonstrates that Qss growth is energetically optimal (12). A more recent study
performed numerical simulations of the impact of a phage T7 infection on its E. coli host,
yielding very interesting insights into the time course of the metabolic demands of a viral
infection (13).
To further explore the energetic requirements of viral synthesis, we made careful estimates of
the energetic costs for two viruses with very different characteristics, namely the T4 phage and
the influenza A virus. T4 phage is a double-stranded DNA (dsDNA) virus with a 169 kb genome
that infects E. coli. The influenza virus is a negative-sense, single-stranded RNA virus (-ssRNA)
with a segmented genome that is 10.6 kb in total length. The influenza virus is a eukaryotic virus
infecting various animals, with an average burst size of 6000 (14). Similar to many other dsDNA
viruses, T4 phage infections yield a relatively modest burst size, with the majority of T4 phages
resulting in a burst size of approximately 200 (15). To determine the energetic demand of
2
viruses on their hosts, the cost estimate for building a single virus has to be multiplied by the
viral burst size and placed in the context of the host's energy budget during the viral infection.
Concretely, the costs associated with building a virus can be broken down into the following
processes that are common to the life-cycles of many viruses: 1) viral entry 2) intracellular
transport, 3) genome replication,4) transcription, 5) translation, 6) assembly and genome
packaging, and 7) exit. Detailed estimates for all of these costs are provided in the SI. Our
strategy was to examine each of these processes for both viruses in parallel, comparing and
contrasting the energetic burdens of each of the steps in the viral life-cycle.
Results
By estimating the energetic costs of influenza and T4 life-cycles, we show that surprisingly the
cost of synthesizing an influenza virus and a T4 phage are nearly the same (Table 1). The
outcome of the analysis to be discussed in the remainder of the paper is summarized pictorially
in Figure 1 for bacteriophage T4 and Figure 2 for influenza. For both viruses, the energetic cost
of translation outweighs other costs (Table 1, Figures 1, 2, 3), though as we will show at the end
of the paper, since translation scales with the surface area of the viral capsid and replication
scales as the volume of the virus, for double-stranded DNA phages larger than a critical size,
the replication cost outpaces the translation cost. Our results will be provided in terms of two
different energetic cost definitions described in detail as part of SI sections II-IV.
To briefly summarize, in our first definition of energetic cost, termed direct cost or 𝐸!, we will
our second definition, termed total cost or 𝐸!, we not only account for the direct costs, but also
for the opportunity cost of building blocks, 𝐸!, required during viral synthesis (SI Figure 1A,
steps 1-4; SI sections II-IV); hence, 𝐸!= 𝐸!+𝐸!. We define the opportunity cost of a building
only account for hydrolysis of ATP molecules (and equivalent molecules, such as GTP) required
during viral synthesis. This definition will include costs such as those incurred during the
synthesis and polymerization of building blocks (SI Figure 1A, steps 3 and 4; SI section II). In
block as the number of ATP molecules that could have been generated had the building block
not been synthesized. The full definition and derivation of these two cost components can be
found in the SI (SI sections II-IV, SI table 5, SI Figure 1, SI Figure 2).
The distinction between these two different energetic cost definitions is that under the direct cost
definition, we attribute energy only to the hydrolysis of ATP-equivalent molecules, whereas
3
under the total cost definition, we also attribute an energetic cost to the building blocks that are
usurped from the host during viral synthesis. Both energy definitions have physical significance.
For example, the direct cost definition only accounts for ATP (and ATP-equivalent) hydrolysis
events, thereby giving us insight into heat production and power consumption of a viral infection.
The total cost definition, on the other hand, is aligned with traditional energetic cost estimates
made from growth experiments in chemostats and allows for a clear comparison between the
cost of an infection and the cost of a cell. To help the reader discern between opportunity and
direct costs, we will signify the former in units of ^P and the latter in the units of ~P. When
reporting total cost estimates, we will simply use P to signify the sum of opportunity and direct
costs.
Moreover, in formulating our estimates, we will generally estimate the cost of a certain viral
process for a single virus, and then multiply this cost by the viral burst size to determine the
infection cost of a given process. Subscript v will denote the cost estimates made for a single
virus, and the subscript i will refer to a cost estimate made for an infection. We relegate the
energetic cost estimates for all viral process to the SI sections V-XI.
The direct, opportunity and total costs of T4 and Influenza. To get a sense for the numbers,
here we provide order-of-magnitude estimates of both the costs of translation and replication
and refer the interested reader to the SI sections II-XI for full details. As detailed in the SI Tables
1 and 2, both T4 and influenza are comprised of about 106 amino acids. We can estimate the
total cost of translation by appealing to a few simple facts. First, the average opportunity cost
per amino acid is about 30 ^P. Second, the average direct cost to produce amino acids from
precursor metabolites is 2 ~P. Finally, each polypeptide bond incurs a direct cost of 4 ~P. We
can see that the total cost of an amino acid is approximately 36 P (30 ^P + 6 ~P). As a result,
the translational cost of an influenza virus and a T4 phage both fall between 107 to 108 P (Table
1).
The cost of viral replication can be approximated in a similar fashion: we have to consider that
the T4 genome is comprised of roughly 4 x105 DNA bases and that the influenza genome is
composed of an order of magnitude fewer RNA bases (≈ 104). The total costs of a DNA
SI Table 5). As a result of T4's longer genome length, its total cost of replication (≈107 P) is
nucleotide and an RNA nucleotide, including the opportunity costs as well as the direct costs of
synthesis and polymerization, are approximately 50 P (SI section II-IX, SI Figure 1, SI Figure 2,
4
kBT on average) (17).
x 1010 P, respectively (SI sections V-XI, Table 1, Figure 1, Figure 3). The total cost of a T4
about an order of magnitude higher than that of an influenza genome (Table 1, Figure 1, Figure
2, SI section VII).
The direct, opportunity and total cost estimates of different viral processes during T4 and
influenza infections are summarized in Figures 1-3 and Table 1. The overall cost of a T4
infection is obtained by summing the costs of replication (𝐸!"#/! ), transcription (𝐸!"/! ),
translation (𝐸!"/! ), and genome packaging (𝐸!"#$/!) required during the infection (SI sections V-
XI, Table 1, Figure 1, Figure 3). These costs together amount to ≈3 x 109 ~P, 8 x 109 ^P, and 1
infection is also equivalent to the aerobic respiration of ≈4 x 108 glucose molecules by E. coli
(26 P per glucose, (16)). Alternatively, it is equivalent to ≈2 x 1011 kBT (assuming 1 ATP = 20
Similarly, the cost of an influenza infection is obtained by adding up the costs of entry (𝐸!"#$%),
intracellular transport (𝐸!"#$%&'/!), replication (𝐸!"#/!), transcription (𝐸!"/!), translation (𝐸!"/!),
and exit (𝐸!"#$/!) required during the infection (SI sections V-XI, Table 1, Figure 2, Figure 3).
These processes have a cumulative cost of ≈8 x 1010 ~P, 5 x 1011 ^P and 6 x 1011 P,
≈2 x 1010 glucose molecules by a eukaryotic cell (32 P per glucose). It is also equivalent to
≈1013 kBT. It is interesting to note that for both viral infections the opportunity cost components
The direct cost of a T4 phage infection is therefore only ≈3% of the direct cost of an influenza
infection. Similarly, the total cost of a T4 phage infection is only ≈2% of the total cost of an
respectively. The sum of costs in an influenza infection is equivalent to the aerobic respiration of
are the dominant component of the total costs.
influenza infection even though individually a T4 phage and an influenza virus have comparable
energetic costs. To contextualize these numbers, the host energy budget (or the host energetic
cost, depending on the viewpoint of a virus versus a cell) during the infection has to be taken
into account.
The total cost of a cell is experimentally tractable through growth experiments in chemostats, in
which cultures are maintained at a constant growth rate. The number of glucose molecules
taken up per cell per unit time can be determined. The number of glucose molecules can then
be converted to an energetic supply by assuming typical conversion rates of 26 P or 32 P per
5
glucose molecule depending on the organism (16). This energetic cost estimate will be a total
cost estimate because not all glucose molecules taken up by the cell are fully metabolized to
carbon dioxide and water to generate ATPs. During the cellular life-cycle, the cell has to double
its number of building blocks prior to division, and to do so, a fraction of glucose molecules
taken up is diverted away from energy production towards biosynthesis pathways. Hence,
cellular energetic cost estimates that are derived from chemostat experiments are total cost
estimates because they report on the combined opportunity and direct costs of a cell (SI section
II-IV).
Based on chemostat growth experiments (18), the total cost of a bacterium and a mammalian
cell with volumes of 1 𝜇𝑚! and 2000 𝜇𝑚!, respectively, are ≈3 x 1010 P and ≈5 x 1013 P, during
of E. coli during its 30-minute doubling time is by considering the dry weight of E. coli (≈0.6 pg)
BNID 100649), an E. coli is composed of ≈2 x 1010 carbons, supplied from ≈3 x 109 glucose
E. coli, this is equivalent to a total cost of ≈7 x 1010 P, which is similar to the number obtained
infection, 𝐸!/!, to the total cost of the host during the infection, 𝐸!/!. For the T4 infection with a
burst size of 200 virions, 𝐸!/! ≈1 x 1010 P (Table 1) and 𝐸!/!≈3 x 1010 P, therefore the
fractional cost of the T4 infection is ≈0.3. Interestingly, a calorimetric study of a marine microbial
despite its larger burst size (6000 virions) and higher 𝐸!/! (≈6 x 1011 P) has a fractional cost of
from chemostat growth experiments (18)(SI section XII).
Moreover, we estimate the fractional cost of a viral infection as the ratio of total cost of an
community demonstrated that 25% of the heat released by microbes is due to phage activity
(20) – an observation that resonates well with our estimate. In contrast, the influenza infection
the course of their viral infections (SI section XII). A simpler estimate for arriving at the total cost
((19), BNID 100089). Given that about half of the cell's dry weight is comprised of carbon ((19),
molecules, since each glucose contributes 6 carbons. With the 26 P per glucose conversion for
just 0.01.
In our estimates for heat production and power consumption of a viral infection, we will not
include the total cost of an infection as it contains the opportunity costs; by definition, these
opportunity costs do not represent direct expenditure of ATP-equivalent molecules and
therefore do not substantially contribute to heat production. In contrast, direct cost estimates
capture only the number of ATP-equivalent molecules hydrolyzed during an infection (SI section
II).
6
T4 infection has a direct cost of 3 x 109 ~P (Table 1). Assuming ATP hydrolysis generates -30
kJ/mole, the heat generated during a T4 infection is approximately 0.1 nJ. An influenza infection
with a direct cost of 7 x 1010 ~P generates 4 nJ. While influenza infection results in an order of
magnitude more heat, the average rate of heat production or the power of T4 and influenza
infections are surprisingly very similar. In half an hour, the T4 infection results in the hydrolysis
of ATP-equivalent molecules at an average rate of 1 x 106 ~P per second. In half a day, an
influenza infection has an average ATP-hydrolysis rate of 2 x 106 ~P per second, which is nearly
the same rate as that of a T4 infection. Put in terms of the more familiar units of Watts, the
power of both viral infections is on the order of 100 fW.
Generalizing viral energetics for double-stranded DNA viruses. While we have concluded
that for the influenza virus and the T4 phage the translational cost outweighs the replication
cost, the ratio of these two costs varies according to the dimensions of a virus. In the case of T4
and influenza, these two viruses had comparable dimensions and consequently were comprised
of a similar number of amino acids (SI Tables 1 and 2). However, due to the diminishing surface
area to volume ratio of a spherical object as it grows in size, the ratio of translational cost to
replication cost also diminishes with increasing radius of a spherical capsid. This simple rule
governs not just nucleotide or amino acid composition of a virus, but more fundamentally, it
governs the elemental composition of viruses with spherical-like geometries (21).
The full derivation of replication and translational cost estimates as a function of viral capsid
inner radius, 𝑟, can be found in the SI section XIII. From these expressions, it is clear that the
translational cost of a virus scales with 𝑟!, whereas the replication cost scales with 𝑟! (Figure
estimates, 𝑟!"#$!!"# (Figure 4, SI section XIII). For the direct cost estimates, the critical radius,
𝑟!"#$!!"#, is 42 nm. Interestingly, a survey of structural diversity encompassing 2,600 viruses
4). The critical radius at which replication will outweigh translation in cost is 59 nm for total cost
inhabiting the world's oceans reveals that the average outer capsid radius is 28 nm (22), which
is much smaller than the predicted critical radii (Figure 4). As such, for the majority of viruses,
we predict translation is the dominant cost of a viral infection.
Furthermore, we provide genome replication to translation cost ratios for about 30 different
double-stranded phages (SI Table 3, Figure 4). While we have omitted calculations for the virus
tails, they can be simply treated as hollow cylinders and will further decrease the expected
7
replication to translation cost ratio for the tailed viruses. Although we have calculated these
ratios for this select group of viruses, similar principles can be applied to modeling the
energetics of other viral groups.
Forces of evolution operating on viral genomes. Inspired by efforts to consider the
evolutionary implications of the cost of a gene to cells of different sizes (18, 23), we were
curious whether similar considerations might be in play in the context of viruses. For example,
we asked which evolutionary forces are prominently operating on neutral genetic elements that
are incorporated into viral genomes, either by horizontal gene transfer, gene duplication or other
similar types of events. We further asked whether the viral size is a parameter of interest in the
tug of war between different forces of evolution. We will address these topics by assuming that
the host lives in an energy-limited environment and that the viral infection, consistent with our
findings for T4, consumes a substantial portion of the host energy budget. By making these
assumptions, we are able to treat the energetic cost of a genetic element as a fitness cost.
For a genetic element to remain in the population, regardless of whether it is beneficial or not, it
must face the consequences of genetic drift which scales with the viral effective population size,
advantage and disadvantage, respectively (Figure 5B). For a genetic element within a viral
genome that is non-transcribed and non-translated (Figure 5C), only the energetic cost of its
replication poses a selective disadvantage. Assuming the genetic element provides no benefit to
𝑁!, as 𝑁!!!. We follow the treatment of Lynch and Marinov who argue that the net selective
advantage of a genetic element is 𝑠!= 𝑠!− 𝑠! , where 𝑠! and 𝑠! denote the selective
the virus (𝑠!=0), the net selective advantage can be stated as 𝑠!= −𝑠!, the absolute value of
which must be much greater than 𝑁!!! for selection to operate effectively. Following Lynch and
element's selection coefficient, 𝑠!, is proportional to its fractional energetic cost, 𝐸! (Figure 5C).
In the case of a non-transcribed genetic element, 𝐸!= !!"#/!!!
, where 𝐸!"#/! corresponds to its
replication cost and 𝐸! is the sum of all costs of a virus (Figure 5C).
Given that replication cost scales as 𝑟! the effects of selection relative to genetic drift could be
Marinov and others (23, 24), we make the simplifying assumption that a neutral genetic
different for viruses of different sizes. Consider Virus A, having a radius that is two times larger
than that of Virus B (Figure 5D). Because both viruses are assumed to have radii larger than the
critical radius, we imagine the scenario in which the cost of genome replication is the dominant
8
cost of synthesizing these viruses. The fractional cost of a genetic element in the smaller virus,
𝐸!_!"#$% ! is then equal to 8𝐸!_!"#$% !, where 𝐸!_!"#$% ! is the fractional cost of the genetic element
in the larger virus. This is because the length of the genome is proportional to 𝑟! , and
consequently, 𝐸! is inversely proportional to 𝑟! (Figure 5D).
Figure 5E and SI Table 4 provide 𝐸! estimates for genetic elements of different lengths (1 –
𝐸! are approximately equal to their replication costs. However, for 𝐸! values in Figure 5E and SI
Table 4, we provide more precise estimates, treating 𝐸! as the sum of both the replication cost
bacterial gene, the direct and total costs of its replication per virus, 𝐸!"#/!, are 3 x 104 ~P and 9
and the translational cost of a virus. The cost of replicating a double-stranded genetic element
can be obtained from SI Eq. 3. For a 1 kb element, which is about the average length of a
10,000 base pairs) within 30 dsDNA viruses. To illustrate the effect of scaling in the example
provided above, we made the simplifying assumption that the viruses are large enough that their
x 104 P, respectively. Both direct and total cost estimates indicate that the strength of selection
acting on a 1 kb, non-transcribed element ranges from 2 x 10-2 - 7 x 10-6 (SI Table 4, Figure 5E)
when considering viruses with radii ranging from ~20 nm to 400 nm. The difference between
direct and total estimates of selection strength is minimal within this range of capsid radii and
continues to diminish as the capsids grow in size.
To examine whether selection or genetic drift will decide the fate of a genetic element we need
to assess each virus's effective population size. This is difficult because the effective population
size of most viruses is unknown and subject to great variability due to several environmental
factors (25). The current effective population size estimates regarding HIV, influenza, dengue,
and measles fall within 101 to 105 (25-27). Based on the wide range of variation in these
effective population sizes, it is difficult to make conclusive statements. It is, however, apparent
that the strength of selection on neutral genetic elements is a non-linear function of the viral
capsid radius and becomes much weaker as viruses get larger (Figure 5E). In fact, for giant
viruses (with outer radius, R > 200 nm), assuming an 𝑁!!! = 10-5, genetic drift could overpower
population. For the majority of viruses (R = 28 ± 6.5 nm, (22)), however, selection is likely to be
selection, allowing for the persistence of neutral elements of lengths 100 bp or shorter in the
the dominant force and drift may only play a role for genetic elements that are just a few base
pairs long (Figure 5E, SI Table 4).
9
Discussion
There have been several experiments that imply a viral infection requires a significant portion of
the host energy budget (5, 6, 8, 10, 28-30). Following these experimental hints, we enumerated
the energetic requirements of two very different viruses on the basis of their life-cycles, and
thereby estimated the energetic burdens of these viral infections on the host cells. According to
our total cost estimates, a T4 infection with a burst size of 200 will consume a significant portion
(about 30%) of the host energy supply. This result, demonstrating a significant fraction of the
host energy used by an infection, supports the experimental findings that the T4 burst size is
correlated positively with the host growth rate (2, 6). It also lends further credence to the
hypothesis that auxiliary metabolic genes within phage genomes are not just evolutionary
accidents; rather, they have come to serve a functional role in boosting the host's metabolic
capacity, which translates into larger viral burst sizes (8, 9, 30, 31). These calculations make it
all the more interesting to develop high-precision, single-cell calorimetry techniques to monitor
energy usage during viral infections. Perhaps the most promising support for T4's cost estimate
is the observation that the maximum T4 burst size is 1,000 virions (15). Using the total cost to
make new viruses, at a burst size of 1,000, the viral infection would consume 170% of the host
energy supply, consistent with the observed apparent upper limits on burst size.
While there are several fascinating studies that explore the link between the host metabolism
and phage infections (8, 11-13), similar studies focusing on viruses of multicellular eukaryotes
are largely lacking. To that end, we chose to estimate the energetic cost of a representative
virus for this category, namely, the influenza virus. The influenza virus and T4 phage are
functionally and evolutionarily very different viruses. Yet, surprisingly, they have a very similar
per-virus cost, regardless of whether the total or the direct cost estimates are being considered.
This is primarily due to the fact that they have a similar translational cost, which dominates all
other costs. And, their comparable cost of translation is due to the fact that these viruses have
similar dimensions and are both composed of about a million amino acids. Perhaps even more
surprising is that both viral infections have very similar average power consumptions, on the
order of 100 fW, despite their different durations.
Even with its higher burst size, an influenza infection has a total cost that is just 1% of the total
cost of a eukaryotic cell. This is because a typical eukaryotic cell is estimated to have much
higher energy supply than a typical bacterium under the same growth conditions. So far in our
estimates, we do not account for the possible inefficiencies at various stages of the viral
10
infection, which may drain more of the host energy than we estimated. Specifically, burst sizes
are typically reported from plaque assays, which count the number of infectious virions that
create plaques. However, we don't have a good estimate for the number of non-infectious
viruses that arise from faulty genome replication, transcription, or viral assembly, for example.
This point is especially important when considering RNA-based viruses such as influenza or
HIV, which have higher mutation rates (10-4-10-6 mutations per base pair per generation; (32))
compared to dsDNA viruses such as T4 (10-6-10-8 mutations per base pair per generation;
((32)). As a result of these higher error rates, RNA-based viruses may have greater hidden
costs associated with aborted viral synthesis or a greater fraction of faulty and non-infectious
virions.
Second, even infectious viruses cannot all be guaranteed to enter the lytic cycle upon infecting
a host cell. In support of these statements is the finding that only about 50% of PBCV-1 viral
progeny are infectious (5). In fact, only 10% of influenza-infected host cells have been shown to
generate infectious virions (33), demonstrating the cumulative inefficiency of an influenza
infection. Hence, counting plaques to measure viral burst sizes may be analogous to making
estimates of the human population by counting only individuals who have children. As such,
single-cell studies of viral infection could provide a detailed breakdown of inefficiencies at
various steps of the viral life-cycle and enable more exact cost estimates. We further explore
other factors related to the fractional cost of influenza and T4 infections in the SI section XIV.
Finally, there is a great need for estimates of the effective population sizes of different viruses
within their natural environments. With current effective population size estimates for viruses it
appears that selection likely determines the fate of genetic elements for the majority of viruses,
which have on average 28 nm radii (22) (Figure 5E, SI Table 4). However, for larger viruses (R
> 200 nm), the diminishing, fractional cost of a gene may enable the interference of genetic drift
to the extent that neutral genetic elements could persist in the viral population. The result of
such a phenomenon could be genome expansions in the form of gene duplication events,
cooption of previously noncoding, horizontally transferred elements into functional genes and
regulatory domains, and perhaps, even a trend towards greater autonomy over large
evolutionary time-scales. This effect may explain the unusual number of duplication events in
the genomes of giant viruses such as that of the Mimivirus (34, 35). Perhaps this effect has also
allowed enough genomic expansion and novelty for certain large viruses to jump the barrier
between obligate entities and self-replicating organisms.
11
Acknowledgements
We are grateful to David Baltimore, Forest Rohwer, Thierry Mora, Aleksandra Walczak, Ry
Young, David Van Valen, Georgi Marinov, Elsa Birch, Yinon Bar-On and Ty Roach for their
many insightful recommendations. This study was supported by the National Science
Foundation Graduate Research Fellowship (grant no. DGE‐1144469), The John Templeton
Foundation (Boundaries of Life Initiative, grant ID 51250), the National Institute of Health's
Maximizing Investigator's Research Award (grant no. RFA-GM-17-002), the National Institute of
Health's Exceptional Unconventional Research Enabling Knowledge Acceleration (grant no.
R01- GM098465), and the National Science Foundation (grant no. NSF PHY11-25915) through
the 2015 Cellular Evolution course at the Kavli Institute for Theoretical Physics.
12
Citations
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
20.
Kutter E & Sulakvelidze A (2004) Bacteriophages: biology and applications (CRC Press).
Hadas H, Einav M, Fishov I, & Zaritsky A (1997) Bacteriophage T4 development
depends on the physiology of its host Escherichia coli. Microbiology 143(1):179-185.
Ahuka-Mundeke S, et al. (2010) Full-length genome sequence of a simian
immunodeficiency virus (SIV) infecting a captive agile mangabey (Cercocebus agilis) is
closely related to SIVrcm infecting wild red-capped mangabeys (Cercocebus torquatus)
in Cameroon. Journal of General Virology 91(12):2959-2964.
Middelboe M (2000) Bacterial growth rate and marine virus–host dynamics. Microbial
Ecology 40(2):114-124.
Van Etten JL, Burbank DE, Xia Y, & Meints RH (1983) Growth cycle of a virus, PBCV-1,
that infects Chlorella-like algae. Virology 126(1):117-125.
Golec P, Karczewska-Golec J, Łoś M, & Węgrzyn G (2014) Bacteriophage T4 can
produce progeny virions in extremely slowly growing Escherichia coli host: comparison
of a mathematical model with the experimental data. FEMS microbiology letters
351(2):156-161.
Rosenwasser S, Ziv C, van Creveld SG, & Vardi A (2016) Virocell Metabolism: Metabolic
Innovations During Host–Virus Interactions in the Ocean. Trends in Microbiology
24(10):821-832.
Lindell D, et al. (2007) Genome-wide expression dynamics of a marine virus and host
reveal features of co-evolution. Nature 449(7158):83-86.
Thai M, et al. (2015) MYC-induced reprogramming of glutamine catabolism supports
optimal virus replication. Nature communications 6.
Chang C-W, Li H-C, Hsu C-F, Chang C-Y, & Lo S-Y (2009) Increased ATP generation in
the host cell is required for efficient vaccinia virus production. Journal of biomedical
science 16(1):1.
Roux S, et al. (2016) Ecogenomics and potential biogeochemical impacts of globally
abundant ocean viruses. Nature.
Kim H & Yin J (2004) Energy‐efficient growth of phage Qβ in Escherichia coli.
Biotechnology and bioengineering 88(2):148-156.
Birch EW, Ruggero NA, & Covert MW (2012) Determining host metabolic limitations on
viral replication via integrated modeling and experimental perturbation. PLoS Comput
Biol 8(10):e1002746.
Stray SJ & Air GM (2001) Apoptosis by influenza viruses correlates with efficiency of
viral mRNA synthesis. Virus research 77(1):3-17.
Delbrück M (1945) The burst size distribution in the growth of bacterial viruses
(bacteriophages). Journal of bacteriology 50(2):131.
Kaleta C, Schäuble S, Rinas U, & Schuster S (2013) Metabolic costs of amino acid and
protein production in Escherichia coli. Biotechnology journal 8(9):1105-1114.
Phillips R, Kondev J, Theriot J, & Garcia H (2012) Physical biology of the cell (Garland
Science).
Lynch M & Marinov GK (2015) The bioenergetic costs of a gene. Proceedings of the
National Academy of Sciences 112(51):15690-15695.
19. 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(suppl
1):D750-D753.
Djamali E, Nulton JD, Turner PJ, Rohwer F, & Salamon P (2012) Heat output by marine
microbial and viral communities.
13
Jover LF, Effler TC, Buchan A, Wilhelm SW, & Weitz JS (2014) The elemental
composition of virus particles: implications for marine biogeochemical cycles. Nature
Reviews Microbiology 12(7):519-528.
Brum JR, Schenck RO, & Sullivan MB (2013) Global morphological analysis of marine
viruses shows minimal regional variation and dominance of non-tailed viruses. The ISME
journal 7(9):1738-1751.
23. Wagner A (2005) Energy constraints on the evolution of gene expression. Molecular
engineer. Biotechnology journal 5(7):686-694.
Anantharaman K, et al. (2014) Sulfur oxidation genes in diverse deep-sea viruses.
Science 344(6185):757-760.
Roux A (2014) Reaching a consensus on the mechanism of dynamin? F1000prime
reports 6.
Lauring AS, Frydman J, & Andino R (2013) The role of mutational robustness in RNA
virus evolution. Nature Reviews Microbiology 11(5):327-336.
Brooke CB, et al. (2013) Most influenza a virions fail to express at least one essential
viral protein. Journal of virology 87(6):3155-3162.
Raoult D, et al. (2004) The 1.2-megabase genome sequence of Mimivirus. Science
306(5700):1344-1350.
Suhre K (2005) Gene and genome duplication in Acanthamoeba polyphaga Mimivirus.
Journal of virology 79(22):14095-14101.
Shepherd CM, et al. (2006) VIPERdb: a relational database for structural virology.
Nucleic acids research 34(suppl 1):D386-D389.
21.
22.
24.
25.
26.
27.
28.
30.
31.
32.
33.
34.
35.
36.
biology and evolution 22(6):1365-1374.
Koonin EV (2015) Energetics and population genetics at the root of eukaryotic cellular
and genomic complexity. Proceedings of the National Academy of Sciences
112(52):15777-15778.
Charlesworth B (2009) Effective population size and patterns of molecular evolution and
variation. Nature Reviews Genetics 10(3):195-205.
Novitsky V, Wang R, Lagakos S, & Essex M (2010) HIV-1 subtype C phylodynamics in
the global epidemic. Viruses 2(1):33-54.
Bedford T, Cobey S, & Pascual M (2011) Strength and tempo of selection revealed in
viral gene genealogies. BMC Evolutionary Biology 11(1):1.
Van Etten JL, Graves MV, Müller DG, Boland W, & Delaroque N (2002)
Phycodnaviridae–large DNA algal viruses. Archives of virology 147(8):1479-1516.
29. Maynard ND, Gutschow MV, Birch EW, & Covert MW (2010) The virus as metabolic
14
Table 1. The direct, opportunity and total energetic costs of viral processes for T4 and influenza.
The T4 infection costs are estimated based on an average burst size of 200, and the influenza
infection costs are based on an average burst size of 6000. Direct costs shown represent the
number of phosphate bonds directly hydrolyzed during the viral lifecycle (~P), whereas the total
costs include both direct costs as well as opportunity costs (^P) incurred during the viral life-
cycle (P). Empty cells correspond to viral processes that did not result in an energetic cost or
were non-applicable to the given virus. See SI sections V-XI.
n
o
i
t
a
c
i
l
p
e
R
n
o
i
t
p
i
r
c
s
n
a
r
T
4 x106 7 x105
3 x105 7 x104
9 x108
108
y
r
t
n
E
l
a
r
i
V
-
-
-
2 x109 4 x108 103
107
7 x105
8 x105 2 x105
2 x109
5 x109
2 x107
108
109
106
106
3 x105
3 x109 3 x108
-
-
-
-
-
-
-
i
g
n
g
a
k
c
a
P
r
a
u
l
l
l
e
c
a
r
t
n
I
t
r
o
p
s
n
a
r
T
t
i
x
E
l
a
r
i
V
n
o
i
t
l
a
s
n
a
r
T
3 x105
-
-
7 x106
-
103
2 x106
7 x107
-
-
107
109
m
u
S
107
107
3 x109
-
-
-
-
-
3 x105
6x106
1010
6 x1010 8 x1010
-
-
-
-
-
-
3 x107
4 x107
3 x107
5 x107
9 x107
-
6 x109
8 x109
2 x1011 3 x1011 5 x1011
-
4 x107
6 x107
-
103
4 x107
6 x107
108
7 x107
-
-
8 x109
1010
6 x109 2 x109 103
-
6 x106 2 x1011 4 x1011 6 x1011
T4
flu
T4
flu
T4
flu
T4
flu
T4
flu
T4
flu
Per
Virion
Per
Infection
Per
Virion
Per
Infection
Per
Virion
t
s
o
c
t
c
e
r
i
d
)
P
~
(
t
s
o
c
y
t
i
n
u
t
r
o
p
p
o
)
P
^
(
t
s
o
c
l
)
P
a
(
t
o
t
Per
Infection
Figure 1. The energetics of a T4 phage infection. The direct, opportunity and total costs of viral
processes are denoted and can be distinguished by their units (~P, ^P and P, respectively).
The energetic requirements of transcription (step 3), translation (step 4), genome replication
(step 5), and genome packaging (step 7) are shown. See SI sections V-XI and Table 1. In this
15
intracellular
figure, when the opportunity cost component is left out, it can be assumed that it is equal to
zero.
Figure 2. The energetics of an influenza infection. The direct, opportunity and total costs of viral
processes are denoted and can be distinguished by their units (~P, ^P and P, respectively). The
energetic requirements of viral entry (steps 2,3),
transport (steps 4,5,9),
transcription (step 6), translation (step 7), genome replication (step 8) and viral exit (step 10) are
shown. See SI sections V-XI and Table 1. In this figure, when the opportunity cost component is
left out, it can be assumed that it is equal to zero.
Figure 3. A breakdown of the direct cost (top) and the total cost (bottom) of various viral
processes during T4 (left) and influenza (right) viral infections (normalized to the sum of all costs
during an infection, as shown in the center of each pie chart). The direct cost of a T4 phage
infection is approximately 3 x 109 ~P (top) while the total cost is 1010 P (bottom). The direct and
total costs of an influenza infection are approximately ~8 x 1010 ~P and 6 x 1011 P, respectively.
Numbers are rounded to the nearest percent, and viral processes costing below 0.5% of the
infection's cost are not shown. See SI sections V-XI for energetic cost estimates for viral entry,
intracellular transport, transcription, viral assembly, and viral exit.
Figure 4. Generalizing viral energetics. A plot of the genome replication (𝐸!"#) to translational
cost (𝐸!") ratio as a function of the virus inner radius, r. The plot uses the geometric parameters
of viruses shown in SI Table 3, all of which are double-stranded DNA viruses with icosahedral
geometries. The predicted numbers of amino acids and nucleotides are derived in SI Table 3..
Cost ratios are shown for both direct and total cost estimates. All viruses shown infect bacteria
except Sputnik, which is a satellite virus of the giant Mimivirus. We have zoomed in on viruses
Sputnik (r = 22 nm), P22 (r = 27.5 nm), T7 (r = 27.5 nm), HK97 (r = 30 nm), and Epsilon15 (r =
31.2 nm). The capsid structures for these representative viruses were obtained from the
VIPERdb (36) and image sizes were scaled based on radii shown in SI Table 3 to accurately
represent the relative sizes of each capsid. The critical radii for the total cost (𝑟!"#$!!"#) and the
direct cost (𝑟!"#$!!"#) estimates are shown. We have also included the mean (𝑟!"#$= 25 nm) and
standard deviation (gray vertical box, ±6.5 nm) of viral capsid inner radii from 2,600 viruses
collected by the Tara Oceans Expeditions (22). Note, here we have subtracted the mean capsid
thickness (3 nm) from the mean capsid radius reported by Brum et al. to arrive at the mean
inner capsid radius.
16
a genetic element within Virus A and Virus B genomes. The fractional cost of a genetic element
Figure 5. Evolutionary forces acting on genetic elements within viral genomes. A) Schematic of
a virus as a spherical object, with an inner radius, r, an outer radius, R, and a capsid thickness,
t. The capsid is composed of viral proteins, while the inner volume holds the viral genome. B)
Positive and negative selective forces (𝑠! and 𝑠!) at a tug of war with the force of genetic drift,
which scales as 𝑁!!!, where 𝑁! is the viral effective population size. C) A schematic of a
genetic element within a viral genome. It is assumed to be non-functional (𝑠! = 0) and non-
transcribed, resulting in 𝑠! = 𝑠! = 𝐸!, where 𝑠! corresponds to the net selection coefficient and
𝐸! corresponds to the fractional cost of a genetic element. D) The evolutionary forces acting on
in Virus B, 𝐸!_!"#$%&, is 8 times higher than the fractional cost of the same element in Virus A,
𝐸!_!"#$%&. Note, Virus A has twice the radius of Virus B, and therefore its genome is 8 times
Both viruses are assumed to have radii greater than critical radii, 𝑟!"#$!!"# and 𝑟!"#$!!"#. E) Log10
𝐸! estimates for non-transcribed and neutral genetic elements of different lengths (1 – 10,000
Table 4; viruses with R > 50 nm are hypothetical dsDNA viruses). Log10 𝐸! estimates derived
estimates, which is not visible in this figure, see SI Table 4). Assuming 𝑁! = 105, the region
(vertical dashed line, 28 nm) and standard deviation (gray vertical box, ±6.5 nm) of viral capsid
base pairs) within the context of 30 dsDNA viruses ranging from ~20 nm to 400 nm in radius (SI
longer than that of Virus B (schematically represented by the number of genetic segments).
from both direct and total cost estimates are included (there is minimal difference between these
above the horizontal dashed line represents a selection-dominated regime, and the region
below it represents a drift-dominated regime. For comparison, we have included the mean
radii obtained from 2600 viruses collected during the Tara Oceans Expeditions (22).
17
Figure 1.
18
Figure 2.
19
Figure 3.
20
Figure 4.
21
Figure 5.
22
|
1001.0284 | 2 | 1001 | 2010-01-06T04:09:03 | Evidence of Conformational Changes in Adsorbed Lysozyme Molecule on Silver Colloids | [
"physics.bio-ph"
] | In this article, we discuss metal-protein interactions in the Ag-lysozyme complex by spectroscopic measurements. The analysis of the variation in relative intensities of SERS bands reveal the orientation and the change in conformation of the protein molecules on the Ag surface with time. The interaction kinetics of metal-protein complexes has been analyzed over a period of three hours via both Raman and absorption measurements. Our analysis indicates that the Ag nanoparticles most likely interact with Trp-123 which is in close proximity to Phe-34 of the lysozyme molecule. | physics.bio-ph | physics | Evidence of Conformational Changes in Adsorbed Lysozyme
Molecule on Silver Colloids
Goutam Chandra1 , Kalyan S. Ghosh2 , Swagata Dasgupta2 and Anushree Roy1
1Department of Physics, Indian Institute of Technology, Kharagpur 721302, India
2Department of Chemistry, Indian Institute of Technology, Kharagpur 721302, India
Abstract
In this article, we discuss metal-protein interactions in the Ag-lysozyme complex by spectro-
scopic measurements. The analysis of the variation in relative intensities of SERS bands reveal the
orientation and the change in conformation of the protein molecules on the Ag surface with time.
The interaction kinetics of metal-protein complexes has been analyzed over a period of three hours
via both Raman and absorption measurements. Our analysis indicates that the Ag nanoparticles
most likely interact with Trp-123 which is in close proximity to Phe-34 of the lysozyme molecule.
Keywords : Lysozyme, Ag colloids, SERS, optical absorption
0
1
0
2
n
a
J
6
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
4
8
2
0
.
1
0
0
1
:
v
i
X
r
a
1
I.
INTRODUCTION
Metal nanoparticles are finding increasing use in nanomedicine for surface interactions
with an array of proteins and/or small molecules. Metal nanoparticles tagged to biomacro-
molecules are oftentimes used to identify specific antibody-antigen binding sites in cells
and tissues [1, 2]. Surface Enhanced Raman scattering (SERS) is an extremely sensitive
technique for monitoring the adsorption of species of very low concentration and for char-
acterizing the structure and orientation of the adsorbed species on the rough metal surface
or metal colloidal particles [3]. Nonresonance SERS is primarily sensitive to the species at-
tached to the metal surface or present within a few of the metal-dielectric interface as well
as to the orientation of the ad-molecule [4]. Thus, it is possible to characterize molecular
rearrangement on the metal surface by SERS mechanism. This implies that conformational
changes occurring in the protein structure or due to denaturation through possible interac-
tions with metal particles can be detected from a detailed analysis of the SERS spectrum.
Modification of the specific interaction sites by covalent or non-covalent methods makes
the possibilities of study enormous. In this regard, the interactions of silver nanoparticles
with several proteins have been investigated. Some of the proteins studied with silver (Ag)
colloids are hemoglobin [5], antirabbit IgG [6], albumin [7] and lysozyme [8–10].
Lysozyme is a small monomeric globular protein consisting of 129 amino acids that has
the ability to disrupt many bacterial functions including their membrane structure [11–14].
The protein has α-helix and β -sheet components with four disulfide bonds. It contains six
tryptophan (Trp) residues, three of them are located at the substrate binding sites, two in
the hydrophobic matrix box, while one is separated from the others [15]. Among them Trp
62 and Trp108 are the most dominant fluorophores, both being located at the substrate
binding sites [16]. Lysozyme action on bacteria has been the focus of a lot of research, but
interaction studies with Ag colloids have not previously revealed the specific interaction sites
on the protein. We have further probed the interactions of lysozyme with Ag nanoparticles
to obtain an insight into the specific residues involved in the interaction. This has been
investigated by spectral methods and new insights have been obtained in the understanding
of how this protein is involved in the interaction with Ag colloids.
Earlier studies on the SERS spectrum of lysozyme molecules adsorbed on Ag electrodes
and hydrosols indicate the proximity of the S-S bond and aromatic amino acid residues of
2
lysozyme molecule on the Ag surface [8–10]. These articles report the orientation of the
aromatic amino acid side-chains located on the surface of the metal electrodes. The role
of an α-helical component of the lysozyme molecule for binding on the metal surface has
been shown [8, 9] and a recent study has reported the conformational changes in lysozyme
molecule with temperature, while adsorbed on a gold nano-patterned SERS substrate [17].
Thus previous studies indicate the possible regions of interaction of the lysozyme molecule
while interacting on the surface of metal colloids. It is necessary to probe the specific region
of binding of the protein molecule to the Ag colloidal surface to permit an understanding
of which regions of the protein would be free to interact with ligands and/or drugs. For
example, the antimicrobial macromolecule formed by a lysozyme-polyphenol complex can
be used as a carrier for phenolic antioxidants. Epigallocatechin gallate (EGCG) [18] and
Triclosan [19] molecules are known to form complexes with lysozyme via Trp-62 and Trp-63
residues. The knowledge of the specific binding site in Ag-lysozyme complex can provide
guidelines to develop nanoparticle tagged lysozyme molecule for drug delivery. It may be
noted that the lysozyme is a potential carrier molecule as a renal-selective drug carrier for
delivery of the angiotensin-converting enzyme (ACE) inhibitor captopril [20].
II. MATERIALS AND METHODS
Synthesis route for preparing stabilized particles: Ag nanoparticles have been syn-
thesized by chemical routes, using sodium borohydride (NaBH4) as reducing agents. Silver
nitrate (AgNO3 ), sodium borohydride (NaBH4) of analytical reagent grade (SRL, India),
were used to prepare the Ag sol. A colloidal silver solution was prepared in deionized wa-
ter following the method described by Creighton et al.
[21]. Essentially, in this chemical
route, AgNO3 is reduced by an excess amount of NaBH4 . 2.2× 10−3 M AgNO3 was added
dropwise to 1 mM NaBH4 at 4 ◦C. Stirring for 20 minutes was necessary to stabilize the
colloidal solution. Later, it was left at room temperature for approximately 1 hour. The
excess NaBH4 evaporated and the remaining solution became transparent yellow in color.
Colloids generated by above method were stable over a considerable period of time (4-6
weeks). However, all our reported results are either with freshly prepared Ag sol or aged
only for a day.
Chicken egg-white lysozyme was obtained from Sigma-Aldrich. In solution, depending on
3
the pH values, the ma jority of lysozyme molecules assume different structural forms. The
protein has a zero net charge at the isoelectric point (pI). Interactions of neutral protein
molecules with water is less favorable than when they are in the charged state. At the pI,
the solubility of proteins in water is low and the repulsion between protein molecules is also
a minimum. These two factors help in adsorption of the molecules on metal surfaces for pH
of the solution close to pI. For lysozyme the value of pI is 10.5. At pH 10.5 the protein is not
soluble in water, which would cause problem in comparing the SERS spectrum and normal
Raman spectrum of lysozyme molecule. At pH 11 the molecule undergoes rapid hydrolysis
[22]. Moreover, the pH of untreated metal sol was between 7.0-7.5. For these reasons we have
maintained the pH of the lysozyme solution in slightly acidic state, such that the effective pH
in the final sol (lysozyme+Ag sol) is close to 7.0-7.5 in all our experiments. 4 mM solution
of lysozyme was prepared in deionized water. For SERS measurement the volume ratio of
Ag sol to protein solution was maintained at 1:1.
SERS spectra were measured using a 488 nm Argon ion laser as an excitation source
using a microRaman spectrometer with 100X ob jective lens. The spectrometer is equipped
with 1200 g/mm holographic grating, a holographic super-notch filter, and a Peltier cooled
CCD detector. The laser power on the samples was 5 mW. The unchanged circular dichro-
ism spectra of laser irradiated and un-irradiated samples eliminate the possibility of dam-
age/conformational changes in the protein structure by laser irradiation under the given
experimental conditions. The data acquisition time for each Raman and SERS spectrum
was 120 sec.
UV-Visible spectra were measured by Spectrascan UV 2600 (Chemito). To study inter-
action kinetics, measurements were carried out immediately after addition of lysozyme in
Ag sol and subsequently after 15 mins, 30 mins, 1 hr, 2hrs and 3 hrs.
For optical absorption measurements the samples were kept in a microcuvette with optical
pathlength 1 cm. For SERS measurements a drop of solution was taken on a glass slide
and allowed to dry.
It took around 10-15 minutes to dry the samples. During kinetic
measurements, to investigate the interaction between metal colloids and protein molecules
in the solution, drops were deposited after the specific time intervals as given above. Thus,
we carried out all measurements after the completion of metal-molecule interaction for the
required time, in ambient conditions. All measurements (SERS and optical absorption) and
analysis have been carried out in triplicate.
4
PyMol [23] was used for visualization of the protein conformation. The accessible surface
area (ASA) of all the amino acid residues in uncomplexed lysozyme are estimated using
NACCESS [24].
III. EFFECT OF ADDED LYSOZYME MOLECULES ON METAL COLLOIDS
A. Analysis of Plasmon Band of Silver Colloids
All biomolecules do not interact with metal colloids in the same way. Some of them result
in agglomeration of the colloidal particles, others do not cause aggregation but modify the
surface charge of the particle [7]. To understand the effect of added lysozyme molecules
on Ag colloids, we study the plasmon band of Ag particles in the sol on adsorption of the
protein molecules. Plasmon band of Ag colloids is shown by black dotted line in Fig. 1
(a). Ag sols absorb light at 398 nm, due to the dipolar surface plasmon of small spherical
particles of Ag. The average particle size of Ag has been estimated to be ∼ 100 A.
Time-dependance of the plasmon band of Ag colloids in the sols is shown in Fig. 1 (a).
Spectra, recorded just after the addition of protein molecules in the metal sols, are shown
by black solid line. Subsequent changes in the spectra with time (after 15 mins., 30 mins, 1
hr, 2 hrs and 3 hrs.) are shown by colored solid lines in the Fig. 1 (a). Addition of lysozyme
to the sol results in (i) decrease in intensity and (ii) broadening with a (iii) red shift of the
plasmon resonance band of Ag particles in the sol. The decrease in relative intensity of
the plasmon absorption band in solutions (ratio of maximum intensity of the plasmon band
of Ag colloids after addition of protein to that of the Ag colloids), is shown in Fig. 1(b).
We observe that the intensity of the plasmon band decays exponentially. The decay time
constant of intensity is estimated to be 60 min in the sol. The observed broadening and
the red shift of the plasmon band indicate formation of bigger particles in the sol by added
lysozyme molecules. Furthermore, the decrease in the intensity of the peak can be explained
by precipitation of metal particles from the sol followed by immediate aggregation.
IV. EFFECT OF METAL COLLOIDS ON ADDED LYSOZYME MOLECULES
Analysis of SERS Spectra
5
Next, we focus on the effect of metal particles on added lysozyme molecules in the sol.
Raman spectra and SERS spectra of 4 mM lysozyme in water and in the Ag sol (just after
addition) are shown by solid lines and dotted lines respectively in Fig. 2. Appearance of the
characteristic feature of the Ag-N vibrational mode at 233 cm−1 [8] in the SERS spectrum
(shown in the inset of Fig. 2) is a clear signature of metal-protein interaction in the sol
via N atom of the ad-molecule. Time variation of SERS spectra of lysozyme in Ag sol
over the spectral range 400–900 cm−1 , 900–1100 cm−1 and 1100–1700 cm−1 are shown in
Fig. 3(a)–(c) The assignments of the prominent bands (indicated by ⋆ marks in Fig. 2)
are summarized in Table I. To study the interaction of the Ag-lysozyme complex, we have
estimated the relative intensity of each feature in the range over 400–1100 cm−1 [Fig. 3(a)
and (b)] with respect to the intensity of the peak at 877 cm−1(vibration of N1H site of Trp
amino acid residues in lysozyme). The methylene δ (CH2 ) vibration at 1442 cm−1 has been
used as reference for the spectral features in the range 1100–1700 cm−1 [Fig. 3(c)]. Both
these peaks (at 877 and 1442 cm−1 ) maintained a constant frequency and were of quite
strong intensities. Invariant Raman shift of a band indicates unchanged chemical bonding
responsible for these particular vibrational modes over experimental duration.
It was interesting to note that the relative intensity of all vibrational bands of the adsorbed
lysozyme molecule do not follow the same trend. A possible explanation for the time-
evolution of the spectra can be the conformational changes of the ad-molecules with time,
which we discuss in detail below.
A. Vibrational modes of amide bands
Vibrations of peptide backbone in proteins, associated with amide I, mainly involve the
C=O stretch with a small contribution of C-N stretching and N-H bending. The same for
the amide III region arises predominantly by an in-plane N-H deformation coupled with a
Cα -N stretch. In lysozyme, amide bands consist of mainly three structural components —
three stretches of α-helix, antiparallel pleated β sheet, in which the polypeptide is hydrogen-
bonded between one region to another via a hairpin turn of the chain and random coils, which
are folded in an irregular way. Lysozyme is a globular protein with distinctly demarcated
α-helix and β -sheet regions. These secondary structural units are connected by random
coils. The positions and intensities of the vibrational modes of amides can be used for an
empirical estimation of these secondary structures in proteins.
SERS spectrum of amide I band appears between 1610–1700 cm−1 . The contributions of
6
the three secondary structures—α-helix, β -sheet and random coil— are expected to occur
over the spectral windows 1665-1680 cm−1 , 1610–1632 cm−1 and 1636–1644 cm−1 , respec-
tively [28]. The amide III band appears over the spectral range between 1230 and 1310
cm−1— the signatures of α-helix, β -sheet and random coils are characterized by spectral
peaks over the range 1280–1320 cm−1 , 1235–1242 cm−1 and 1250–1260 cm−1 , respectively
[29]. The region 1254–1260 cm−1 has been assigned to β -turn [30].
In the measured SERS spectra of lysozyme in Ag sol, the vibrational modes of amide
III appear at 1260 cm−1 , as a shoulder of the strong peak at 1331 cm−1 of Trp (Fig. 3).
Thus, it is difficult to extract the subtle spectral features of amide III band from spectral
analysis. To study the correlation between three secondary structural components of the
adsorbed protein molecule, the broad envelope of the amide I band of each spectrum needs
to be deconvoluted into the contributions from each of them. Such analysis may also provide
information on conformational changes in lysozyme with time upon adsorption on colloidal
surface. The time-variation of the vibrational spectra of the amide I band are shown in Fig.
4 by + signs. The deconvoluted and net fitted spectra are shown in Fig. 4 by dotted and
solid lines. From the analysis of the spectral profile of amide I band we obtain the following:
(i) the spectral feature over the spectral range between 1610 and 1632 cm−1 for β -sheet
is missing in all spectra and (ii) the relative vibrational intensities of random coils (IR ) to
α-helix (Iα ) increases with time (in the inset of Fig. 4).
B. Vibrational modes of amino acid residues
As mentioned before, the sharp SERS spectral line at 877 cm−1 corresponds to the vibra-
tion of N1H site of Trp residues [31]. The frequency of this band depends on the strength
of the H-bond between Trp and other side chain molecule. The same feature is expected to
shift by 8 cm−1 for non-H-bonding Trp derivatives and lowered by 7 cm−1 for H-bonded Trp
derivatives. The single sharp peak at 877 cm−1 in the Raman and SERS spectra indicates
that in the metal-protein complex the Trp-residues do not form any new chemical bond
directly with the Ag surface or with other residues. The variation in relative intensities
of other vibrational modes of Trp are shown in Fig. 5(a). The intensity of ring breathing
mode of Phe-residue at 1004 cm−1 first increases and then decreases with time (Fig. 5(b)).
The change in δ (CH2 ) mode of glutamic acid at 1449 cm−1 (just next to δ (CH2 ) methylene
vibrational mode at 1442 cm−1 ) is shown in Fig. 5(c). In each plot, the solid line is the
guide to the eye.
7
We did not find any clear signature of Asp and His (histidine), two other amino residues
of lysozyme, in the SERS spectra of the molecule.
C. Disulfide bond
There are four pairs of S-S bond in lysozyme. Among these four pairs, two pairs of S-S
bonds are near the α-helix region and two others are in the β -sheet region. The strong SERS
spectrum for vibration of S-S bond appears at 505 cm−1 just after addition of lysozyme in
the Ag sol. However, the intensity of this band decreases with time (Fig. 5(d)). To check
the effect of the reducing environment on the disulfide bond, an additional experiment with
lysozyme and NaBH4 was conducted. The trend in the change of relative intensity of S-S
bond (Fig. 5(e)) indicates that reduction of the disulphide linkage by BH−
4 (note that the
Ag colloidal particles are stabilized by BH−
4 in NaBH4 stabilized sol) is possible.
Discussion
Here we report the specific amino acid residues of lysozyme molecule which are adsorbed
on Ag colloids. Appearance of strong vibrational modes of Trp and Phe residues in the SERS
spectra of lysozyme molecule and the similar trend in the change in relative intensities of
these modes (Fig. 5(a) and Fig. 5(b)) signify that both these residues are close to the
Ag surface and occupying adjacent sites in the adsorbed molecule. A lysozyme molecule
contains six Trp residues (Fig. 6). The accessible surface area (ASA) available for these
residues are shown in Table II. Three (Trp-62, Trp-63 and Trp-123) of the six Trp residues
of a free lysozyme molecule have ASA more than 50 A2 and, hence, are likely to be free to
bind with a substrate. Trp-62 and Trp-63 are exposed at the edge of the active site cleft and
known to interact specifically with a substrate [29]. Trp-123 is close to the c-end of the amino
acid chain of the molecule. Trp-28, -108, -111 have relatively less ASA and are expected to
be burried. Of all three Phe residues (Phe-3, Phe-34 and Phe-38) of lysozyme molecule the
ring of Phe-34 is very close to Trp-123 and its ASA is relatively large compared to that of
the other two [Fig. 6 and Table II]. Thus, it is most likely that the lysozyme molecules are
adsorbed on Ag surface via the red marked region which includes both Trp-123 and Phe-34
residues of the molecule [Fig. 6].
Next we discuss the observed SERS spectra of amide I band. It is to be noted that the
ratio of the secondary structure contents i.e α helix:β sheet:random coil in free lysozyme is
expected to be 0.54:0.15:1 [32]. Dominant spectral features of α-helical component in Fig. 4
indicate that lysozyme is adsorbed on the colloidal Ag surface either in an α-helical region
8
or in a loop with an α-helical element [33]. The increase in the intensity ratio (IR : Iα) in the
inset of Fig. 4 indicates unfolding of the α-helix or the α-helical turn element to a random
coil of adsorbed protein molecule with time. It is interesting to note that the φ,ψ angles of
Trp-123 in lysozyme are -124.9,-2.5 [red arrow in Fig. 6], that is close to the α-helical region
of the Ramachandran plot.
Furthermore, the disulfide bond is expected to be affected by the change in protein
backbone conformation [34]. Of all four S-S bonds, two of them, (S6-S127) and (S30-S115)
are very close to Trp-123. As the turn element, at the arrow in Fig. 6, gets affected due
to the interaction with the Ag surface, S6-S127/S30-S115 bond/bonds may break down
and this explains the decrease in the intensity of S-S vibrational mode with time (Fig.
5(d)). Furthermore, Glu-7 of lysozyme lies very close to the S6-S127 bond with an ASA
corresponding to 82.4 A2 . If the S-S bond breaks down (Fig. 5(d)), the vibrational mode
of Glu-7 is expected to be affected in a similar fashion (Fig. 5(c)). Moreover, it is to be
noted that this particular region of the molecule corresponds to the C-terminal end of the
protein. Hence, this region of the molecule is more flexible and susceptible to changes in the
environment.
Addition of lysozyme in Ag sol results in an immediate agglomeration of the Ag particles.
The initial increase in the relative intensity of vibrational bands of amino-acid residues of
the molecule with time in Fig. 5(a) and Fig. 5(b) suggest re-orientation of the residues to
obtain a low energy configuration from an initial high energy state due to initial favorable
electrostatic interactions during change in conformation. The time constant for the unfold-
ing of the α-helix is also estimated to be 45 min (inset of Fig. 4). If we assume only the local
electrostatic interaction between the adsorbates and metal colloids, all our results suggest
that lysozyme molecules interact with Ag colloids instantaneously; however, it takes longer
time for the protein molecule to be stabilized on the metal surface. In general, irreversible
laws of thermodynamics are involved in adsorption of proteins on metal surfaces [35]. Var-
ious time scales are involved in metal-molecule interactions. The fast steps are sometimes
reversible (can be probed only by time-resolved experimental measurements) and the rela-
tively slow steps (of time-scale varies from seconds to minutes) result in rearrangements of
protein structure by the surface environment and in many cases, can be irreversible [35]. We
believe that the initial metal-protein interaction via agglomeration of the particles involves
the fast process mentioned above and the conformational changes of the lysozyme molecule
9
with time corresponds to the latter process. All above conjectures, obtained from spec-
troscopic measurements, can be further confirmed by studying interaction of Ag-lysozyme
complexes with other ligands that are currently underway, for which the interaction sites of
lysozyme are already established.
As mentioned earlier that there are few reports in which SERS spectra of lysozyme
molecules on Ag colloidal surface have been discussed. Prior knowledge available on the
interaction of the Ag-nanoparticles with lysozyme indicates the α-helical and random coil
components [9]. Our experimental approach identifies the particular Trp residue involved in
the interaction. Besides this, the effect of the interaction on the intensities of SERS bands
of Phe and disulfide bonds confirm the specific site of the protein involved in the complexa-
tion. Conformational changes in the protein molecule on interaction with Ag nanoparticles
indicate partial unfolding of the α-helix constituted of residues 25-35 of which Phe-34 is a
member. Interestingly, it is to be noted that the parts of the protein molecule with β -sheet
comprising of Trp-62 and Trp-63, known to have higher bio-activity, remain unaffected by
metal-protein complexation.
Acknowledgements
AR and SDG thank DRDO, India for financial assistance
10
[1] M. Horisberger, In Preparation of Biological Specimen for Scanning Elecetron Microscopy J.A.
Murphy,G.M. Roomans, Eds. Scanning Electron Microscopy Inc. Chicago, IL, 1984; pp. 315-
336.
[2] J. Roth, In Techniques in Immunocytochesmistry G.R. Bullock, P. Petrusz, Eds. Academic
Press, London, 1983; Vol. 2, pp-216-284.
[3] K. Kneipp, H. Kneipp, I. Itzkan, R.R. Dasari, M.S. Feld, Chem. Rev. 99 (1999) 2957.
[4] C.L. Haynes, A.D. McFarland, R.P. VanDuyne, Anal. Chem. (2005) 338A.
[5] H. Xu, E.J. Bjerneld, L. Kall, M. Borjesson, Phys. Rev. Lett. 83 (1999) 4357.
[6] E.S. Grabbe, R.P. Buck, J. Electroanal. Chem. 308 (1991) 227.
[7] A.M. Ahern, R.L. Garrell, Langmuir 7 (1991) 254.
[8] J. Hu, R.S. Sheng, Z.S. Xu, Y. Zeng, Spectrochim. Acta 51A (1995) 1087.
[9] E. Podstawka, Y. Ozaki, L.M. Proniewicz, Appl. Spectrosc. 58 (2004) 1147.
[10] G.D. Chumanov, R.G. Efremov, I.R. Nabiev, J. Raman Spectrosc. 21 (1990) 43.
[11] A. Pellegrini, U. Thomas, N. Bramaz, S. Klauser, P. Hunziker, R. von Fellenberg, J. Appl.
Microbi. 82 (1997) 372.
[12] H.R. Ibrahim, T. Matsuzaki, T. Aoki, FEBS Lett. 506 (2001) 27.
[13] A. Pellegrini, U. Thomas, P. Wild, E. Schraner, R. von Fellenberg, Microbiol. Res. 155(2)
(2000) 69.
[14] T. Croguennec, F. Nau, D. Molle, Y.L. Graet, G. Brule, Food Chem. 68 (2000) 29.
[15] C.J. Sheng, H.D. Dian, Lysozyme, Shandong Science and Technology Press, 1982, pp. 50-51.
[16] T. Imoto, L.S. Foster, J.A. Ruoley, F. Tanaka, Proc. Natl, Acad. Sci. USA 69 (1972) 1151.
[17] G. Das, F. Mecarini, F. Gentile, F. D. Angelis, HG. M. Kumar, P. Candeloro, C. Liberale, G.
Cuda, E. D. Fabrizio, Biosensors and Bioelectronics 24 (2009) 1693.
[18] K.S. Ghosh, B.K. Sahoo and S. Dasgupta , Chemical Physics Letts. 452 (2008) 193.
[19] Md. Imranul Hoq, K. Mitsuno, Y. Tsujino, T. Aoki, H.R. Ibrahim, International Journal of
Biological Macromolecules 42 (2008) 468.
[20] Jai Prakash, A.M. van Loenen-Weemaes, M. Haas, J.H. Proost, D.K.F. Meijer, F. Moolenaar,
11
K. Poelstra and R.J. Kok, Drug Metabolism and Disposition 33 (2005) 683.
[21] J.A. Creighton, C.G. Blatchford, M. Grant Albrecht, J. Chem. Soc. Faraday Trans. 75 (1979)
790.
[22] R.C. Lord, Nai-Teng Yu, J. Mol. Biol. 50 (1970) 509.
[23] W.L. DeLano, The PyMOL molecular graphics system DeLano scientific, San Carlos, CA,
USA: 2004.http://www.pymol.org
[24] S.J. Hubbard, J.M. Thornton, ’NACCESS’. Computer Program, Department of Biochemistry
and Molecular Biology, University College London; 1993.
[25] M.C. Chen, R.C. Lord, R. Mendelsohn, J. Am. Chem. Soc. 96 (1974) 3038.
[26] C.H. Chuang, Y.T. Chen, J. Raman Spectrosc. 40 (2008) 150.
[27] J.T. Lopez Navarrete, V. Hernandez and F.J. Ramirez, Journal of Molecular Structure 348
(1995) 249.
[28] A.C. Dong, P. Huang, W.S. Caughey, Biochemistry 29 (1990) 3303.
[29] J.L. Lippert, R.M. Lindsay, R. Schultz, Biochimica et Biophys. Acta 599 (1980) 32.
[30] R.C. Lord, Appl. Spectrosc. 31 (1977) 187.
[31] T. Miura, H. Takeuchi, I. Harada, Biochemistry 30 (1991) 6074.
[32] J.L. Lippert, D. Tyminski, P.J. Desmeules, J. Am. Chem. Soc. 98 (1976) 7075.
[33] M. Pal, S. Dasgupta, Proteins: Structure, Function, and Genetics 51 (2003) 591.
[34] C.H. Munro, W.E. Smith, M. Garner, J. Clarkson, P.C. White, Langmuir 11 (1995) 3712.
[35] R. Eckert, S. Jenny, J.K.H. Horber, Cel l Biology International 21 (1997) 707.
12
Figure Captions
Figure 1. (a)Kinetic study of the plasmon band of Ag colloids and (b)Time variation in
relative intensity of the plasmon band of the Ag upon addition of lysozyme in the sol.
Figure 2. Comparison between normal Raman spectrum of 4×10−3M lysozyme in water
(solid lines) and SERS spectrum of the same in the Ag sol (dotted lines). Inset of the figure
shows the characteristic feature of Ag-N vibrational mode.
Figure 3. SERS spectra of Ag-lysozyme complex taken at different time over the range
(a)450-900 cm−1 , (b) 900-1100 cm−1 and (c) 1100-1700 cm−1 .
Figure 4. SERS spectrum (+ signs) and deconvoluted spectra (dashed lines) of the amide
I band. Inset of the figure shows the variation in intensity ratio of random coil to α-helix
components of the protein with time.
Figure 5. Variation in intensity of vibrational bands of (a)Trp residues (b)Phe residues
(c)glutamic acid (d)S-S bonds of adsorbed lysozyme in Ag sol and (e)S-S bonds of lysozyme
in BH4 environment (without metal colloid) with time respectively.
Figure 6. Cartoon of lysozyme molecule and its site of interaction with Ag surface (the
region within the red mark. The α-helix element which unfolds upon interaction with metal
surface is shown by arrow.
13
Table Captions
Table I. Assignment of vibrational bands for lysozyme SERS spectra.
Table II. Calculated accessible surface area (ASA) for Trp, Phe and Glu residues in a
lysozyme molecule.
14
)
2.5
s
t
i
n
2.0
u
.
b
r
1.5
a
(
e
c
1.0
n
a
b
0.5
r
o
s
b
0.0
A
(a)
Ag_NaBH4
Ag + Lys (0 min)
After 15 mins
After 30 mins
After 1 hr
After 2 hrs
After 3 hrs
0.75
y
0.70
t
i
s
0.65
n
e
0.60
t
n
I
0.55
e
v
0.50
i
t
a
0.45
l
e
R
0.40
(b)
400
600
1000
800
Wavelength (nm)
0
40
80
120
Time (min)
160
200
FIG. 1: Chandra et al
1600
1200
800
400
0
)
s
t
i
n
u
.
b
r
A
(
y
t
i
s
n
e
t
n
I
200
280
240
Raman shift (cm-1)
3000
)
s
2500
t
i
n
u
2000
.
b
r
1500
A
(
y
1000
t
i
s
n
e
t
n
I
500
0
400
600
1200
800
Raman shift (cm-1)
1600
FIG. 2: Chandra et al
15
3000
)
s
t
2500
i
n
u
2000
.
b
r
1500
A
(
y
1000
t
i
s
n
e
t
n
I
500
0
Ag+Lys(just after mixing)
after 15 min
after 30 min
after 60 min
after 120 min
after 180 min
(a)
500
600
800
700
Raman shift (cm-1)
900
3000
)
s
2500
t
i
n
u
2000
.
b
r
1500
A
(
y
1000
t
i
s
n
e
t
n
I
500
0
900
Ag+Lys(just after mixing)
after 15 min
after 30 min
after 60 min
after 120 min
after 180 min
(b)
950
1050
1000
Raman shift (cm-1)
1100
12000
)
s
t
i
n
9000
u
.
b
r
6000
A
(
y
t
i
3000
s
n
e
t
n
I
Ag+Lys(just after mixing)
after 15 min
after 30 min
after 60 min
after 120 min
after 180 min
(c)
0
1100 1200 1300 1400 1500 1600 1700
Raman shift (cm-1)
FIG. 3: Chandra et al
16
1.6
R
1.2
0.8
0.4
0
120 180
60
Time (min)
(180 min)
(120 min)
2000
1000
0
1000
0
1000
(60 min)
0
1000
0
1000
0
1000
(30 min)
(15 min)
(0 min)
)
s
t
i
n
u
.
b
r
A
(
y
t
i
s
n
e
t
n
I
0
1600
1660
1620
1640
-1
Raman shift (cm
)
1680
1700
FIG. 4: Chandra et al
17
Trp (a)
1015 cm-1
0
0
30 60 90 120
1356 cm-1
30 60 90 120
1568 cm-1
760 cm-1
30 60 90 120
1330 cm-1
30 60 90 120
1542 cm-1
0.4
0.2
0.0
1.2
0.8
0.4
0.0
1.2
0.8
0.4
0.0
30 60 90 120
0
Time (min)
30 60 90 120
1.2
0.8
0.4
0.0
0.8
0.4
0.0
0.8
0.4
0.0
o
i
t
a
r
e
v
i
t
a
l
e
R
0
0
0
1004 cm-1
Phe (b)
0
30
90
60
Time (min)
120
505 cm-1
S-S band (d)
o
i
t
a
r
e
v
i
t
a
l
e
R
1.2
0.8
0.4
0.0
1.2
o
i
t
0.8
a
r
e
v
i
0.4
t
a
l
e
R
0.0
Glu (c)
1.2 1450 cm-1
0.9
0.6
0.3
0.0
o
i
t
a
r
e
v
i
t
a
l
e
R
-0.3
0
120
30
90
60
Time (min)
1.2
505 cm-1
S-S band (e)
o
i
t
a
r
e
v
i
t
a
l
e
R
0.8
0.4
0.0
0
30
60
90
Time (min)
120
0
30
60
90
Time (min)
120
FIG. 5: Chandra et al
18
FIG. 6: Chandra et al
19
TABLE I: Chandra et al
Assignment
peak (cm −1 ) Reference
S-S
Phe
Tyr
Trp - indole breathing (W18)
Trp- N1H site vibration (W17)
Phe - ring breathing
Trp - ring breathing
Trp- indole ring (C-H deformation)
Trp- W7 wagging
δ(CH2 )
Glu-δ(CH2 )
Trp -indole ring stretching
Trp -W2 mode
Phe/Tyr
Amide - I
505
620
644
763
877
1004
1015
1338
1356
1442
1449
1540
1568
1598
1644
[8, 22]
[25]
[25]
[26]
[26]
[8]
[8]
[26]
[26]
[25]
[27]
[26]
[26]
[8]
[8]
20
TABLE II: Chandra et al
Trp sequence ASA (A2 ) Phe-sequence ASA (A2 )
Trp-28
0.12
Phe-3
Trp-62
120.57
Phe-34
Trp-63
Trp-108
Trp-111
Trp-123
48.01
10.59
12.69
55.5
Phe-38
Glu-7
Glu-35
17.82
64.02
15.33
82.44
34.16
21
|
1805.06429 | 1 | 1805 | 2018-05-16T17:11:12 | The propagation of active-passive interfaces in bacterial swarms | [
"physics.bio-ph"
] | Propagating interfaces are ubiquitous in nature, underlying instabilities and pattern formation in biology and material science. Physical principles governing interface growth are well understood in passive settings; however, our understanding of interfaces in active systems is still in its infancy. Here, we study the evolution of an active-passive interface using a model active matter system, bacterial swarms. We use ultra-violet light exposure to create compact domains of passive bacteria within Serratia marcescens swarms, thereby creating interfaces separating motile and immotile cells. Post-exposure, the boundary re-shapes and erodes due to self-emergent collective flows. We demonstrate that the active-passive boundary acts as a diffuse interface with mechanical properties set by the flow. Intriguingly, interfacial velocity couples to local swarm speed and interface curvature, suggesting that an active analogue to classic Gibbs-Thomson-Stefan conditions controls boundary propagation. Our results generalize interface theories to mixing and segregation in active systems with collective flows. | physics.bio-ph | physics | The propagation of active-passive interfaces in bacterial swarms
Alison E. Patteson1,+,∗, Arvind Gopinath2,3 and Paulo E. Arratia1,∗
1Department of Mechanical Engineering & Applied Mechanics,
University of Pennsylvania, Philadelphia, PA 19104.
2Department of Bioengineering, University of California Merced, CA 95340.
3Health Sciences Research Institute, University of California Merced, CA 95340
+Present Address - Physics Department, Syracuse University NY 13244.
Propagating interfaces are ubiquitous in nature, underlying instabilities and pattern for-
mation in biology and material science. Physical principles governing interface growth are
well understood in passive settings; however, our understanding of interfaces in active sys-
tems is still in its infancy. Here, we study the evolution of an active-passive interface using a
model active matter system, bacterial swarms. We use ultra-violet light exposure to create
compact domains of passive bacteria within Serratia marcescens swarms, thereby creating
interfaces separating motile and immotile cells. Post-exposure, the boundary re-shapes and
erodes due to self-emergent collective flows. We demonstrate that the active-passive bound-
ary acts as a diffuse interface with mechanical properties set by the flow.
Intriguingly,
interfacial velocity couples to local swarm speed and interface curvature, suggesting that an
active analogue to classic Gibbs-Thomson-Stefan conditions controls boundary propagation.
Our results generalize interface theories to mixing and segregation in active systems with
collective flows.
8
1
0
2
y
a
M
6
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
9
2
4
6
0
.
5
0
8
1
:
v
i
X
r
a
1
Bacteria live and move in an extraordinarily wide range of habitats and can quickly respond
to the presence of other cells and physical boundaries in their environment. For instance, bacteria
swim independently in fluids, but when transferred to surfaces display a collective behavior known
as swarming [1, 2]. Swarming occurs in many gram-negative and gram-positive species and cor-
responds to a hyper-flagellated elongated phenotype [2–4]. Swarming cells self-organize into rapid
collective motions that allow for quick colonization of new environmental niches [1, 5–7]. Swarm-
ing is co-regulated with virulence determinants, inversely-regulated with sessile biofilm formation,
and associated with enhanced antibiotic resistance [7–9]. More broadly, the collective motion of
self-propelling (active) particles [10, 11] is observed in bacterial infections [12], embryogenesis [13],
and wound healing [14] and is an important feature of both prokaryotic [15, 16] and eukaryotic
[17, 18] systems.
Ecological niches are typically a heterogeneous mix of cells, and internal boundaries can form
separating cells of two difference types. Bacterial swarms coexist symbiotically with other microbes
– assisting in the transport of fungal spores [19] and other bacterial species [20] – or they compete
at sharp boundaries [21, 22]. Boundaries also emerge in cultures of the same bacteria due to
chemotaxis and cell death [23, 24] or the presence of extracellular polymers [25–27], both of which
can induce a swimming-speed dependent phase separation [24, 27]. Segregation of active particles
is not unique to biological settings, arising in synthetic systems such as phoretic particles [28, 29].
In passive bi-phasic systems [30–32] such as melting ice-water mixtures and solidifying alloys,
properties of internal boundaries (such as interface shape and speed) depend on the surface tension,
interfacial energies, and externally imposed flows. In active systems, particle motion can couple to
the presence of boundaries which can lead to anomalous properties in mechanical pressure [33, 34]
and impact collective flows [35, 36]. However, despite the ubiquity of boundaries in living and life-
like materials, boundary stability and motion remain largely unexplored in active non-equilibrium
matter. Identifying boundary conditions that link boundary structure and active motion will help
elucidate a minimal description of actively-driven boundaries.
Here, we focus on the propagation of an interface separating mobile and immobile bacteria in
swarming Serratia marcescens. Serratia are a rod-shaped, gram-negative, opportunistic pathogen
of the Enterobacteriaceae family [3]. We use high-intensity ultra-violet (UV) light to selectively
paralyze and passivate cells in large compact domains within the swarm (Methods). The passive
domain and active swarm interact at the interphase boundary where self-emergent, vortical flows
develop. The interphase is spontaneously reshaped and eroded as passivated bacteria are dislodged
from their neighbors and convected by nearby collective flows.
Intriguingly, the active-passive
2
boundary behaves as a propagating, diffuse elastic interface with speeds that correlate with local
interface curvature and the intensity of the active bacterial flow. Our results suggest the existence
of an active analog to the classic Gibbs-Thomson-Stefan boundary conditions of passive multiphase
systems.
Figure 1a shows snapshots of a Serratia marcescens swarm before and after its exposure to UV
light. The swarm is grown on an agar substrates (Methods 1) and is pictured moving right to left
at a speed of approximately 1 µm/min; the colony edge is marked by a clear precursor fluid film
(white edge in image). Close examination reveals (Fig. 1b) densely-packed rod-shaped cells with
local orientational order resembling a nematic liquid crystal. The individual cells [3] are 1 µm in
diameter and 5-30 µm in length, and the collective swarm edge is estimated to be approximately
a monolayer thick based on previous investigations [6]. In its initial state, the swarm is highly
motile and exhibits long-range collective flows (SI-Movie 1). We use particle image velocimetry
(PIV, Methods 2, SI-§II) to exact the bacterial velocity field v(r, t) (overlaid in color). We find
that the fastest moving cells are approximately 100-400 µm from the leading edge of the colony,
where the average local speed is 28 µm/s (Fig. 1c). The collective flow has a correlation length of
approximately 20 µm and characteristic time of 0.25 seconds (SI-§II, SI-Fig. 1).
When exposed to sufficiently high intensity light (Methods 3, SI-I), a portion of the highly-
motile swarm is quenched and becomes immobile (Fig. 1a, SI-Movie 2). We use standard micro-
scope optics to focus the light from a mercury lamp source to the swarm, selectively damaging and
blocking cell motility [Methods 3, SI-I]. This generates domains of passive bacteria, the shape of
which can be controlled through an aperture. We focus on two aperture geometries: a half-plane
aperture [H] and an octagonal-shaped [O] aperture.
Figures 1 illustrates the swarm's response to UV light using the [O] aperture. Bacteria inside the
exposed region (highlighted octagon) stop moving and are eventually trapped. This is seen in SI-
Movie 2: particles that serve as tracers slow down and eventually stop moving as they are trapped
amongst the passive bacteria. Bacteria outside the exposure remain motile, forming vortices and
jets along the boundary of the trapped cells (SI-Movie 3).
The collective flow of bacteria shapes and erodes the passive domain. When the light source is
switched off, the active unexposed bacteria penetrate the passive region, dislodge, and thereafter
convect passive bacteria away from the interface (SI Movie 4). The interface propagates radially
inward albeit asymmetrically. The swarm's collective motion recovers to its pre-exposure state
(Fig. 1c).
Flow streamlines (Fig. 1d) near the interphase boundary show the rich complex behaviors of
3
individual vortices. We observe vortices convecting in from the bulk, interacting with each other,
and lingering at the surface (highlighted brown vortex, for example). Similar features occur when
the exposure is made using the [H] aperture (SI-Fig. 2, SI-Movie 5); a notable difference for the [H]
aperture is that the absence of imposed corners and the approximate translational invariance along
the exposed edge results in active-passive boundary propagating without large scale curvatures.
Bi-phasic passive systems involving boundary erosion such as melting of ice or alloy solidifica-
tion usually feature a diffuse interfacial region where the two phases mix. We hypothesized that,
similarly, the boundary between the passive (immobile) and active (swarming) bacteria may be
treated as an diffuse interface possessing an intrinsic time dependent thickness, w. We also hy-
pothesized that a suitable (mesoscale level) continuous phase-field order parameters could be used
to define an interface with physically relevant properties. Thus we defined two order parameters
φ (Methods 4, SI-III): (i) φ - based on intensity fluctuations between images of the swarm, which
are related to bacterial density fluctuations, and (ii) φv - based on the bacterial velocity fields.
Each order parameter defines a phase field φ(r, t) that quantifies the local motion throughout
the swarm as shown in Fig. 2: the active phase corresponds to values of +1 and the passive phase,
-1. A transitional interfacial region exists between the two phases; we define the mathematical
interface as the locus of points rs satisfying φ(rs, t) = 0 (Fig. 2a-i & b-i). We find that the results
obtained from using the order parameters φ and φv yield essentially the same description of the
active-passive interface as seen in the phase field maps in Fig. 2a-ii & b-ii; here, we present results
primarily from φ based on intensity fluctuations (see SI-Fig. 4 for φv results).
To examine the boundary thickness, we spatially-average the phase field into one-dimensional
phase profiles, φ∗ (Fig. 2a-iii & b-iii, Methods 5). Physically, we intuit that the one-dimensional,
spatially-averaged phase profiles correspond to a propagating, broadening, and coarsening interface,
whose analytical form corresponds to a hyperbolic tangent.
Indeed, our phase profiles are well
captured using fits to hyperbolic tangent functions (Fig. 2a-iii & b-iii, Methods 5), which yield
the mean interface position d and interface thickness w over time.
The evolution of the mean interface position d depends on the shape of the aperture (Fig. 2a-iv
& b-iv). For the (O) aperture, d(t) provides an area averaged radius of the passive domain that
decreases over time, eventually reducing to zero as the whole domain erodes away; in this case, we
t0 − t (Fig. 2a-iv), with the time at which the passive phase disappears, t0 ≈ 40 s. For
t for the
find d ∼ √
the (H) aperture (Fig. 2b-iv), the mean interface position propagates faster than d ∼ √
duration of the experiment.
The interface thickness w ranges from 4 to 10 µm (∼ bacterial length) over most of the disso-
4
lution process (Fig. 2a-v & b-v). The most significant deviation from this trend is seen for the
[O] aperture at relatively long times (Fig. 2a-v): when the interface width is approximately the
radius of the passive phase (t ≈ 40 s), both w and d exhibit large fluctuations and the thickness
increases dramatically. Our results show that the interface width is a weak function of time (Fig.
2a-v & b-v), consistent with a quasi-steady propagating interface. We do observe isolated bacteria
- singly or in very small pockets - entering into the passive region at distances more than the inter-
face thickness (SI Movie 4). However, these events are rare, resulting in the averaging procedure
yielding a phase-field based interface without overhangs. We interpret w as a correlation length
characterizing the gradient in the density of active bacteria.
The active-passive phase fields based on intensity fluctuations (φ) and bacterial velocity field
(φv) yield qualitatively similar results (SI-III). For instance, the interface positions are the same
(SI-Fig. 4). We hypothesize that wv is greater than w (SI-Fig. 4) because velocity fields vary over
vortex (20 µm) length scales whereas intensity fluctuations vary over bacterial (5 µm) scales.
Next, we examine how the bacterial flow interacts with the interface. Figure 3a highlights
two robust features of the flow near the interface: (i) a gradient in flow perpendicularly across
the interface and (ii) an array of size-varying vortical flows. We find that the flow varies in both
strength and dynamics. As shown in 3a (inset), the square of vorticity (cid:104)ω2(cid:105) increases as one
moves away from the interface; while the flow decay time τ decreases (Methods 6), indicating an
increase in vortical flow lifetimes, a result similar to recent simulations of active and passive sphere
mixtures [36]. Overall, our data suggests a coupling between the interface and flow: the interface
plays a stabilizing role on the collective flow while the gradient in bacterial vorticity marks a flux
of momentum and energy from the swarming bacteria to the interface, energy that can be used to
erode and reform the surface. Consistent with this picture, the interface undergoes displacements
through interactions with many fluctuating vortical flows (Methods 6).
To understand the boundary-flow coupling, we calculate the (quasi-static) structure factor of
the interface (Fig. 3b) and the two-dimensional energy spectrum of the flow (Fig. 3c). We
focus here on experiments using the [H] aperture, which are not limited by finite time dissolution
effects. The static structure factor (cid:104)∆hq2(cid:105) (Methods 6) decays with wavenumber q, punctuated
by peaks at q = 0.15 µm−1 and 0.22 µm−1 (Fig. 3b). For q > 0.4 µm−1, the overall decay scales as
(cid:104)∆hq2(cid:105) ∼ q−2. To correlate these length scales with the energetic features of the flow, we calculate
the two-dimensional time-averaged energy spectra E(q) from the bacteria velocity field (Methods
6) for different distances from the interface Y . As shown in Fig. 3c, E(q) is non-monotonic
with q and depends on Y . This non-monotonic behavior is a unique characteristic of active fluids
5
and is attributed to the injection of energy at the level of the bacteria [37]. Compared to dense
suspensions of swimming bacteria in microfluidic devices [37], our measured spectra exhibit similar
scalings with q as q5/3 and q−8/3 (Fig. 3c) with a peak centered at scales of about 1-2 vortex sizes.
The interface also impacts the flow spectra by shifting the peak to higher q, indicating a shift to
smaller vortices near the interface (Fig. 3c).
For passive fluid-fluid interfaces fluctuating due to white noise, equipartition of energy requires
that (cid:104)∆hq2(cid:105) = (κAq2)−1 with κ being the stiffness, and A the interface area. Equipartition does
not hold in our system. Comparing E(q) to (cid:104)∆hq2(cid:105) (Fig. 3b & c), we find that the peak in
the E(q) (between 0.15-0.2 µm−1; length scale, 15-20 µm) correlates with peaks in (cid:104)∆hq2(cid:105) - a
signature that the interface is forced at length scales of the flow structures (streamers and vortices).
Still, since the scaling (cid:104)∆hq2(cid:105) ∼ q−2 provides a good description over a range of length scales
(Fig. 3b), we use this form to fit the data (Methods 7) and extract an interfacial stiffness κ ≈ 0.7
µm−2. This value is much smaller than the interfacial stiffness of water in air (1.7×107 µm−2) but
similar to phase-separated systems involving colloids [38–40] (0.1-20 µm−2). This stiffness is to be
interpreted as an effective value of the diffuse interface and inherently accommodates variations in
passive bacteria alignment.
Here, the bacterial flow injects energy into the surface at various wave-numbers and frequencies.
One way to qualify the energetic features of the flow is by invoking an effective temperature Teff . In
the absence of flow, Teff can be estimated by following tracers that sample the system for sufficiently
long times [41]. Our swarm features collective flows; Teff therefore is ambiguous. Nonetheless, using
2 µm polystyrene spheres as tracer particles (Methods 7, SI-Fig. 10), we estimate Teff ≈ 2.2 × 105
K, which yields an apparent surface tension κkBTeff ∼ 10 pN/µm.
In the paradigmatic example - melting of ice - the interface between ice and water propagates
with speeds controlled by the temperature field at the interface and heat flux across the interface.
For a stationary ice-water boundary, the Gibbs-Thomson relationship provides the relationship
between the interface curvature C and the temperature (equivalently, chemical potential) while the
Stefan boundary condition constrains the flux of heat across the interface. For slowly propagating
ice-water interfaces, the Gibbs-Thomson relationship requires extension; a linear relationship can
be obtained between the local curvature, local interface velocity and temperature (or chemical
potential) [31, 32].
To test for analogous boundary conditions for active-passive interfaces, we visualized how active
flows extracted and convected passive particles at the boundary (Fig.4a). Based on these obser-
vations (see Fig. 4(a), Methods 8) we propose a simple linear relation that captures the main
6
ingredients of the erosion process relating the interface kinematics to geometry and intensity of
activity,
vint = a1 + a2C + a3X
(1)
with the variable X standing in for the swarm activity. We chose bacterial speed v, vN (normal
component of the bacterial velocity), vT (tangential component), and vorticity ω as possible stand-
ins for X and evaluated the fidelity of the fit of Eqn. 1 to the experimental data (Methods 8,
SI-§V). We gather these data at multiple points along the interface for varying times throughout
the experiment. The data is combined with local measures of the interface velocity and interface
curvature and cast into three-dimensional scatter plots (SI-§ V,SI-Fig. 9). The best collapse of
the data to a plane (p < 0.05) was obtained for vN, as shown in Fig. 4b & 4c (SI-§V, Methods
8). The values of the coefficients are dependent on the aperture geometry. For the half-plane
[H] aperture, we find a1 = 0.4 µm/s, a2 = −2.3 µm2/s and a3 = −2.0. For the octagonal [O]
aperture, the presence of a mean curvature shifts the plane, which we find for vN to be given by
a1 = 0.02 µm/s, a2 = −1.12 µm2/s, and a3 = −0.7. At a qualitative level, the scatter plots in Fig.
4 demonstrate three-way coupling between the interface velocity, bacterial flow, and curvature in a
manner reminiscent of the classical Gibbs-Thomson-Stefan boundary conditions in passive systems.
The interface velocity is negatively correlated with curvature C. Regions of negative curvature move
into the passive phase as passive bacteria are extracted from the surface.
We conclude that dissolution of passive domains within active swarms is a spontaneous self-
organized process governed by short-range interactions at the single bacterial level and long-range
non-local effects due to collective flow. Three features define the active-passive interface during
the erosion process. First, the interphase region acts as a broadening diffuse interface with a well
defined position and thickness. Second, the interface is stiff with elastic constants dependent on
flow intensity; stronger flows are expected to lower the effective stiffness, allowing for faster erosion.
Third, the interface structure is tightly coupled to the statistics of the collective swarming flows.
The interphase boundary stabilizes vortices that form parallel to the surface, enabling sustained
erosion. The net result is that the interface erosion follows a relation bearing similarities to the
classic Gibbs-Thomson-Stefan boundary conditions. These results generalize classical interface
theories that couple mechanics, thermodynamics and kinetics to active living and synthetic systems
with collective flows.
Interfaces separating cells of differing motility is a common motif in natural habitats, including
7
biofilms, where swimmers, spores, and other non-motile phenotypes segregate into different domains
[42]. Starting with light-induced boundaries in a model system of Serratia marcescens swarms, we
identified an active version of Gibbs-Thomson and Stefan boundary conditions. We anticipate
that these physical mechanisms underlie segregation and pattern formation in active biological and
synthetic settings. Natural next steps would be to explore how extracellular polymers - which
are implicated in both single cell swimming [43, 44] and collective biofilm expansion [45] - impact
boundary dynamics in microbial environments.
METHODS
1. Preparation of bacterial swarms and imaging
We use swarming Serratia marcescens cultured on on agar substrates. The agar is prepared by
dissolving 1 w% Bacto Tryptone, 0.5 w% yeast extract, 0.5 w% NaCl, and 0.6 w% Bacto Agar in
ddH2O. This is melted and poured into petri dishes after addition of 2 w% of glucose solution (25
w%). Once the agar cools and solidifies, Serratia marcescens (strain ATCC 274, Manassas, VA)
from frozen glycerol stocks are inoculated on the agar plates and incubated at 34o C. Bacterial
colonies form at the inoculation sites and grow outward on the agar substrate, away from the
inoculation site. Experiments are performed 12-16 hours after inoculation; the bacteria are imaged
with the free surface facing down with an inverted microscope. The bacteria were imaged with an
inverted Nikon microscope Eclipse Ti-U with a Nikon 20x (NA = 0.45) and Nikon 60x (NA = 0.7)
objective. Images were gathered at either 60 or 125 frames per second with a Photron Fastcam
SA1.1 camera.
2. Bacterial velocity fields from PIV
Bacterial velocity fields, v(r, t), were extracted at 3 µm intervals from videos using particle image
velocimetry (PIV, SI-§II). The velocity fields reveal complex transient collective flows at the ex-
panding edge of the colony. Selecting regions of the swarm spanning 100 to 500 µm from the colony
edge - where the swarming behavior appears strongest - we calculated the spatial and temporal
velocity correlation functions, Cr(∆r) and Ct(∆t) using
(cid:68) v(r0) · v(r0 + ∆r)
(cid:69)
,
(2)
Cr(∆r) =
v(r0)2
and
(cid:68) v(t0) · v(t0 + ∆t)
v(t0)2
(cid:69)
.
Ct(∆t) =
8
(3)
Here, the brackets denote and average over reference times t0 and reference positions r0. Pre-
exposure, the swarming behavior does not appear to be biased in any particular direction (the
swarming speeds are much larger than the speed of the advancing colony front); hence, scalar
function defined by (1) and (2) suffice to characterize the flow. From these measurements, we
extract a characteristic vortex size of 20 µm and vortex (residence or lifetime) timescale of 0.24
seconds (SI-§II, SI-Fig. 1).
3. Exposure to high intensity ultraviolet light
To create a passive phase of immobile bacteria, we used standard fluorescence microscope optics
to expose cells to high intensity light. The light source is an unfiltered mercury lamp, which has
a wide-spectrum that includes a significant amount of ultra-violet (UV) light (100-400 nm). The
response of bacteria to light depends on the wavelength, intensity, and duration of the light [46] and
is affected by photosensitizers [46, 47]. High intensity light ≥ 200 mW/cm2 at wavelengths 390-530
nm can lead to increased tumbling, slow swimming, and eventual irreversible paralysis [46, 47]. It
is hypothesized that paralysis may be due to flagellar motor damage caused by photosensitizing
flavins and dyes that are present in growth media such as LB Broth. Experiments [48] with Bacillus
subtilis also suggest that photosensitizers can be used to reversibly or irreversibly affect collective
motility. Cell death occurs from photodynamic action and UV light-induced DNA damage.
Consistent with previous experiments [46–48], we found that for swarming Serratia marcescens,
passivation was not immediate and sometimes reversible; yet for sufficiently large light intensities
and exposure times they were rendered permanently passivated. We selected an exposure time of
60 s and light intensity of I = 370 µW (measured at 535 nm) to reproducibly immobilize Serratia
marcescens (SI-I).
4. Phase-field order parameters identify interface
We use two scalar order parameters, φ and φv, based on a phase-field description of the interface
region between the active (motile) and passive (non-motile) domains to define, identify, locate and
characterize the interface boundary (see also SI-§III).
9
(I) From intensity fluctuations: The first scalar order parameter, φ is computed from intensity
fluctuations, ∆I, defined by the difference in image intensities taken at 0.1 second intervals of the
bacterial swarm (SI-§III-1). Treating the intensity fluctuation as a measure of the fluctuations in
the bacteria density, we define the order parameter at location r by
φ(r, t) =
2∆I(r, t) − ∆IA(t) − ∆IP (t)
∆IA(t) − ∆IP (t)
.
(4)
where ∆IA(t) and ∆IP (t) are the intensity fluctuations of the active and passive phases far from
the interface. Here, −1 ≤ φ ≤ +1, with -1 corresponding to the completely passive phase and 1
corresponding to the completely active phase.
The interfacial region is fuzzy with finite thickness due to the intermingling of active and passive
bacteria. Following classical phase-field approaches, we mathematically define the interface location
as the locus of points satisfying φ = 0.
(II) From velocity fields through PIV: The second order parameter φv is based on the velocity
fields calculated from PIV (SI-§III-2, SI-Fig. 3). We define
φv(r, t) =
2v2(r, t) − v2
A − v2
v2
P
A − v2
P
(5)
A and v2
where v2
P are velocities of the active and passive phases far from the interface and r is the
position of velocity vectors sampled at 3 µm intervals. Again we have the bounds −1 ≤ φv ≤ +1
with the interface location defined as the locus of points satisfying φv = 0.
5. Position and thickness of the diffuse interface
The boundary separating the active and passive phases resembles a propagating and broadening
interface. To analyze this behavior, we fit the order parameter profiles to the classical form for
a quasi-equilibrium diffuse, moving interface. We use the order parameter fields extracted from
intensity fluctuations (Fig. 2) and PIV data (SI-§III) to obtain averaged one-dimensional phase
profiles.
For experiments with the [H] aperture, we choose cartesian coordinates (x, y) with the y = 0
line aligned with the edge of the exposure. For times post-exposure, fields are averaged in the x
direction to obtain φ∗(y, t). Data from experiments then is fit to
φ∗ = tanh
y − d(t)
w(t)
and φ∗
v = tanh
(cid:35)
.
y − dv(t)
wv(t)
10
(6)
(cid:34)
(cid:34)
(cid:34)
(cid:35)
,
(cid:35)
,
Here d(t) and dv(t) denote the interface location and w(t) and wv(t) denote the diffuse interface
thickness.
In experiments using the [O] exposure, we define a polar co-ordinate system with center at
the center of the passive phase and obtain an azimuthally-averaged, radially-dependent value of
φ∗(r, t). Reduced data is then fit to the forms
φ∗ = tanh
r − d(t)
w(t)
and φ∗
v = tanh
r − dv(t)
wv(t)
(7)
(cid:34)
(cid:35)
where now d(t) is the averaged radius of the passive domain. Since r = 0 is the center of the
passive region, equation (7) will exhibit significant errors for times close to dissolution (d ≤ w).
The length-scale over which φ varies sharply, w(t), yields the density correlation length char-
acterizing the penetration of active bacteria into the passive phase. The length-scale over which
φv varies sharply, wv(t), yields the momentum penetration length - the length over which passive
bacteria are pushed around by the active bacteria without being completely dislodged.
6.
Interface-flow coupling
(i) Height fluctuations To compare the structural features of the collective flow with the fea-
tures of the interface, we calculated the static structure factor of the interface after suitably averag-
ing fluctuations on the order of bacterial lengths. Focusing on experiments with the [H] aperture,
we compute the height fluctuation ∆h(t, x) = h(t, x) − d(t), where h(t, x) is the interface position
interpolated as a function of the arc-length coordinate x and d(t) is the mean interface position.
Next, the wavenumber dependent Fourier modes of the height fluctuations, ∆hq(t), was obtained
as
(cid:90) Lx
0
1
Lx
∆hq(t) =
∆h(x, t) e−iqx dx.
(8)
Here, the wave number is q = nπ/Lx where Lx ≈ 200 µm). We varies q from 0.015 µm−1, which
corresponds to the system size, to 1.0 µm−1, which corresponds to approximately 3 µm (∼ half
a bacterial length). The static structure factor was determined as the temporal-average of the
Fourier mode magnitude square, (cid:104)∆hq2(cid:105).
11
The energy transfer to the surface comes from the flux of momentum flowing into the passive
domain from the active domain as active bacteria invade and erode the interface.
(ii) Energy spectra The two-dimensional energy spectra of the flow is defined here through the
PIV-velocity fields as
(cid:90)
E(q) =
q
2π
dR e−iq·R(cid:104)v(t, r0) · v(t, r0 + R)(cid:105),
(9)
where the brackets here denote averages over r0. For the near-interface case, we calculate E(q)
within an [80 × 80] µm2 area, adjacent to the interface (Y ranging from 0-80 µm; average distance
40 µm). For the far from interface case, we calculate E(q) within an [100 × 100] µm2 area, 100
µm+ from the interface (Y ranging from 80-180 µm; average distance 130 µm). We varied q from
approximately 0.03 µm−1 to 1.0 µm−1, corresponding to approximately the size of the region (80
µm) and to the bacterial velocity field spacing (3 µm), respectively. For the [H]-aperture results
(Fig. 3), the interface position is a flat propagating line, and Y is the distance normal to this line.
(iii) Temporal correlation functions We calculated the autocorrelation function of the bound-
ary position (using the [H]-aperture), defined here as
Cint(∆t) =
(cid:104)∆h(t0)∆h(t0 + ∆t)(cid:105)
(cid:104)∆h(t0)2(cid:105)
,
(10)
where the correlation is averaged over reference locations and reference times t0. The data is
fit to a single decaying exponential, yielding a characteristic decay time τ of 16 s (SI-Fig. 9).
For characterize the bacterial flow, we measure the normalized spatial correlation of the velocity
director v (from PIV) as a function of time ∆t and the distance normal to the interface Y , as given
by
Cflow(∆t, Y ) =
(cid:104)v(t0, Y )v(t0 + ∆t, Y )(cid:105)
(cid:104)v(t0, Y )2(cid:105)
(11)
where the correlation is averaged over reference locations and reference times t0. The flow corre-
lations are also fit to a single decaying exponential to determine characteristic flow time scales as
a function of Y (SI-Fig. 9).
12
7.
Interfacial stiffness and effective temperature
To extract an effective interfacial stiffness κ, we fit the structure factor in Fig. 3b to (κAq2)−1,
assuming that the interfacial area is h × LX , where h = 1 µm is the cell width (since the swarm is
approximately a monolayer thick [6]) and LX is the length of the observed interface, 200 µm.
To estimate an effective temperature of the swarm, we use two micron polystyrene spheres as
probes of the swarming flows in the active region of the colony (SI-Fig. 10). The polystyrene
particles are cleaned by centrifugation and then suspended in a buffer solution (67 mM NaCl
aqueous solution) with a small amount of surfactant (Tween 20, 0.03% by volume). A small
aliquot of this particle solution (20 µL) is gently pipetted unto the bacterial colony in a location
where expanding colony front meets the agar. After the polystyrene particles are introduced, we
allow the colony to settle for 5 minutes before imaging. We do not observe any change in the
behavior of the swarming bacteria due to the addition of these particles at 0.8 % area fraction. We
track the particle positions over time using standard particle tracking techniques. We define the
(time-averaged) particle speed as the particles displacement (in two dimensions) over a 1 second
time interval, allowing tracers to sample multiple vortex structures (characteristic lifetimes ∼ 0.1
second). The particle speed distributions seem to follow a 2D Maxwell-Boltmann distribution
(SI-Fig. 10b), p(v) = vm(kBTeff)−1exp(−mv2/2kBTeff), where m is the mass of the polystyrene
particle, kB is the Boltzmann constant, and Teff ≈ 2.2×105 K, approximately 700 times the thermal
temperature (293 K). This effective temperature is to be interpreted as a mixture temperature
purely due to the energy in the swarming collective flows.
8. Effects of curvature and flow on interface speed
Does an approximate linear relationship between the interface speed, curvature, and a characteristic
of the bacterial flow exist for small propagation speeds? Such a relationship may be viewed then as
an active analogue of the classical Gibbs-Thomson-Stefan boundary conditions for passive systems.
Swarming provides the impetus for the motion and keeps the active-passive interface always out of
equilibrium. We can directly gauge the intensity of the flow and its activity via measurements such
as PIV or particle tracking methods. This is preferable to using temperature or chemical potential
that cannot be measured directly and exact forms of which are a matter of debate [11, 33, 34].
To test for correlations between the interface speed, curvature and flow speeds we assume
that two independent processes act on the interface leading to erosion and remodeling. The first
13
arises from self-propulsion based interactions at the single bacteria level and is proportional to
density of active bacteria ρA. The second acts at longer spatiotemporal scales and accounts for
hydrodynamic interactions and collective flow. Let vint(r) be the interface velocity at location
r on the interface; the (mesoscale) normal at this location is n(r, t), the (mesoscale) tangent is
t(r, t), and the (mesoscale, collective) bacterial velocity field is v. Let Q be the alignment tensor
that quantifies the orientation of the jammed passive bacteria in the neighborhood of r. Ignoring
quadratic and higher order terms (in ∇n, ∇t, ∇v and ∇Q), we write the general linear form
vintn ≈ ρAA1 + ρAA2 (∇ · n) + ρA(n · Q · n) v +
ρA(n · Q · t) v + ρA(t · Q · t) v + ρA(t · Q · n) v +
ρA(v · Q · n + n · Q · v + t · Q · v + v · Q · t)
(12)
with A1 and A2 being functionals of Q, n and t. The first term on the right hand side reflects
local density driven self-propulsion effects present even in the absence of curvature or collective
motion. The second curvature term reflects variations in the erosion rate due to geometric effects
and bacteria preference to reside near the surface. The other terms describe erosion driven by the
collective flow and include normal and tangential velocity contributions. Equation (12) as written
cannot be used directly to interpret our data as we do not measure or visualize the alignment field
of bacteria in the passive region.
To gain insight into the form of this linear relationship, we studied the trajectories of 2 µm
polystyrene tracer particles trapped in the passive phase to identify the sequence of mechanistic
events leading to their eventual extraction (Fig.4a). From observations on dozens of tracers, we find
that erosion occurs due to initially trapped passive particles being dislodged from their neighbors,
then moving parallel to the surface as they are sheared by active bacteria (SI-Movie 6), and even-
tually escaping and leaving that passive domain due to normal streaming flows that occur between
vortices at the interface (Fig.4a). Guided by these observations of tracer particles escaping from
the passive phase (Fig.4a), we further deduce that anisotropic caging effects result in tangential
and normal swarm velocities to affect the dislodgment of the passive particles differently. At the
14
simplest level this incorporates the non-isotropic nature of Q. Incorporating these ideas, we write
vint ≈ ρA(α1 + α2 C + α3(v·n) + α4(v·t)).
(13)
Deviations from this form can be attributed to non-linear effects involving curvature and flow,
neglect of alignment effects in the passive phase and density variations in the active phase. While
we expect α1 (and thus a1 = ρAα1) to be zero in the absence of flow and curvature, we retain this
as a fitting constant as higher order terms and the colony front velocity may lead to a net interface
propagation.
[1] Harshey, R. M. Bacterial motility on a surface: many ways to a common goal. Annual Reviews in
Microbiology 57, 249-273 (2003).
[2] Kearns, D. B. A field guide to bacterial swarming motility. Nature Reviews Microbiology 8, 634-644
(2010).
[3] Alberti, L. & Harshey, R. M. Differentiation of Serratia marcescens 274 into swimmer and swarmer
cells. Journal of bacteriology 172, 4322-4328 (1990).
[4] Harshey, R. M. & Matsuyama T. Dimorphic transition in Escherichia coli and Salmonella typhimurium:
Surface-induced differentiation into hyperflagellate swarmer cells. Proc. Natl Acad. Sci. USA 91, 8631-
8635 (1994).
[5] Steager, E. B., Kim, C. B. & Kim, M. J. Dynamics of pattern formation in bacterial swarms. Physics
of Fluids 20, 073601 (2011).
[6] Darnton, N. C., Turner, L., Rojevsky, S. & Berg, H. C. Dynamics of bacterial swarms. Biophysical
journal 98, 2082-2090 (2010).
[7] Verstraeten, N. et al. Living on a surface: swarming and biofilm formation. Trends in Microbiology 16,
496-506 (2008).
[8] Kim, W., Killam, T., Sood, V. & Surette, M.G. Swarm-cell differentiation in Salmonella enterica
serovar Typhimurium results in elevated resistance to multiple antibiotics. J. Bacteriol. 185, 3111-
3117 (2003).
[9] Butler, M.T., Wang, Q. & Harshey, R. M. Cell density and mobility protect swarming bacteria against
antibiotics. Proc. Natl Acad. Sci. USA 107, 3776-3781 (2010).
[10] Marchetti, M. et al. Hydrodynamics of soft active matter. Reviews of Modern Physics 85, 1143 (2013).
[11] Patteson, A. E., Gopinath, A. & Arratia. P. E. Active Colloids in Complex Fluids. Current Opinion
Colloids and Interfacial Science 21, 86-96 (2016).
[12] Hall-Stoodley, L., Costerton, J.W. & Stoodley, P. Bacterial biofilms: From the natural environment to
infections diseases. Nature Reviews Microbiology 2, 95-108 (2004).
15
[13] Scarpa, E. & Mayor, R. Collective cell migration in development. J. Cell Biol. 212, 143-155 (2016).
[14] Vitorino, P. & Meyer, T. Modular control of endothelial sheet migration. Genes Dev. 22, 3268 (2008).
[15] Ben-Jacob, E., Cohen, I. & Gutnick, D.L. Cooperative organization of bacterial colonies: from genotype
to morphotype. Annu Rev Microbiol 52, 779-806 (1998).
[16] Koch, D.L. & Subramanian, G. Collective Hydrodynamics of Swimming Microorganisms: Living Fluids.
Annual Review of Fluid Mechanics 43, 637-59 (2011).
[17] Krishna Vedula, S., Ravasio, A., Lim, C.T. & Ladoux, B. Collective Cell Migration: A Mechanistic
Perspective. Physiology 28,370-379 (2013).
[18] Mehes, E. & Vicsek, S. Collective motion of cells: from experiments to models. Integr. Biol. 6, 831-854
(2014).
[19] Ingham, C. J., Kalisman, O., Finkelshtein, A. & Ben-Jacob, E. Mutually facilitated dispersal between
the nonmotile fungus Aspergillus fumigatus and the swarming bacterium Paenibacillus vortex. Proc.
Natl Acad. Sci. USA 108, 19731-19736 (2011).
[20] Finkelshtein, A., Roth, D., Jacob, E. B. & Ingham, C. J. Bacterial swarms recruit cargo bacteria to
pave the way in toxic environments. MBio 6, e00074-00015 (2015).
[21] Gibbs, K.A., Urbanowski, M.L. & Greenberg, E.P. Genetic determinants of self identity and social
recognition in bacteria. Science 321, 256-259 (2008).
[22] Budding, A.E. et al. The Dienes Phenomenon: Competition and Territoriality in Swarming Proteus
mirabilis. Proc. Natl Acad. Sci. USA 107, 13626-13630 (2010).
[23] Woodward, D.E. et al. Spatio-Temporal Patterns Generated by Salmonella typhimurium. Biophysical
Journal 68, 4052-4057 (1995).
[24] Cates, M.E., Marenduzzo, D., Paonabarraga, I. & Tailleur, J., Arrested phase separation in reproducing
bacteria creates a generic route to pattern formation. Proc. Natl Acad. Sci. USA 107, 11715-11720
(2010).
[25] Grant, M.A.A., Waclaw, B., Allen, R.J. & Cicuta, P. The role of mechanical forces in the planar-to-bulk
transition in growing Escherichia coli microcolonies. Interface 11, 20140400 (2014).
[26] Ghosh, P., Mondal, J., Ben-Jacob, E. & Levine, H. Mechanically-driven phase separation in a growing
bacterial colony. Proc. Natl Acad. Sci. USA 112, E2166-E2173 (2014).
[27] Schwarz-Linek, J., et al. Phase separation and rotor self-assembly in active particle suspensions. Proc.
Natl Acad. Sci. USA 109, 4052-4057 (2012).
[28] Palacci, J. et al. Living crystals of light-activated colloidal surfers. Science 339, 936-940 (2013).
[29] Takatori, S., De Dier, R., Vermant, J. & Brady, J. Acoustic trapping of active matter. Nature Com-
munications 7, 10694 (2015).
[30] Bray, A.J. Coarsening dynamics of phase-separating systems. Philos Trans A Math Phys Eng Sci. 15,
781-91 (2003).
[31] Langer, J.S. Instabilities and pattern formation in crystal growth. Rev. Mod. Phys. 52, 1 (1980).
[32] Back, J.M., McCue, S.W., & Moroney, T.J.. Including nonequilibrium interface kinetics in a continuum
16
model for melting nanoscaled particles. Sci. Rep. 4, 7066 (2014).
[33] Takatori, S. C., Yan, W. & Brady, J.F. Swim pressure: stress generation in active matter. Physical
Review Letters 113, 028103 (2014).
[34] Solon, A.P., et al. Pressure is not a state function for generic active fluids. Nature Physics 11, 673-678
(2015).
[35] Lushi, E., Wioland, H. & Goldstein, R.E. Fluid flows created by swimming bacteria drive self-
organization in confined suspensions. Proc. Natl Acad. Sci. USA 111, 9733-9738 (2014).
[36] Wysocki, A., Winkler, R. G. & Gompper, G. Propagating interfaces in mixtures of active and passive
Brownian particles. New Journal of Physics 18, 12 (2016).
[37] Wensink, H. H. et al., Meso-scale turbulence in living fluids. Proc. Natl Acad. Sci. USA 109, 14308-
14313 (2012).
[38] Hernandez-Guzman, J. & Weeks, E. R. The equilibrium intrinsic crystal-liquid interface of colloids.
Proc. Natl Acad. Sci. USA 106, 15198-15202 (2009).
[39] Gasser, U., et al. Real-space imaging of nucleation and growth in colloidal crystallization. Science 292,
258-262 (2001).
[40] Aarts, D. G., Schmidt, M. & Lekkerkerker, H. N. Direct visual observation of thermal capillary waves.
Science 304, 847-850 (2004).
[41] Patteson, A.E., Gopinath, A., Purohit, P.K. & Arratia, P.E. Particle diffusion in active fluids is
non-monotonic in size. Soft matter 12, 2365-2372 (2016).
[42] Chai, L, Vlamakis, H. & Kolter, R. Extracellular signal regulation of cell differentiation in biofilms.
MRS Bull. 36, 374-379 (2011).
[43] Patteson, A.E., Gopinath, A., Goulian, M. & Arratia P.E. Running and tumbling with E. coli in
polymeric solutions. Scientific Reports 5, 15761 (2015).
[44] Berg, H.C. & Turner, L. Movement of microorganisms in viscous environments. Nature 278, 349-351
(1979).
[45] Gloag, E.S., et al. Self-organization of bacterial biofilms is facilitated by extracellular DNA. Proc. Natl
Acad. Sci. USA 110, 11541-11546 (2013).
[46] Taylor, B.L. & Koshland, D.E. Intrinsic and Extrinsic Light Responses of Salmonella typhimurium and
Escherichia coli. Journal of bacteriology 123, 557 (1975).
[47] Conley, M.P. & Berg, H.C. Chemical Modification of Streptococcus Flagellar Motors. Journal of
bacteriology 158, 832 (1984).
[48] Lu, S. et al. Loss of Collective Motion in Swarming Bacteria Undergoing Stress. Physical Review Letters
111, 208101 (2013).
ACKNOWLEDGEMENTS
17
We thank Edward Steager, Elizabeth Hunter, Somayeh Fahardi, Mark Goulian, Prashant Purohit,
Julia Yeomans, and James Sethna for fruitful discussions. This work was supported by NSF-
CBET-1437482. A. E. Patteson was supported by an NSF Graduate Research Fellowship.
AUTHOR INFORMATION
The authors declare no competing financial interests.
∗Correspondence and requests for materials should be addressed to A.E.P. ([email protected]) and
P.E.A. ([email protected]).
18
FIG. 1. Creation and dissolution of an active-passive interphase boundary in a bacterial swarm
(a) Snapshot of the expanding edge of a swarm of Serratia marcescens on agar; the colony front exhibits
a precursor fluid film (∼ 5 µm; white curve), moving from right to left. The swarm shows long-range
collective flows, with strong velocity fields (PIV; overlaid color). A large domain of passive, immobile
bacteria is created by exposing a region of the swarm to high intensity ultraviolet (UV) light (highlighted
octagon). An interphase boundary forms between the passivated and active bacteria. When the light source
is switched off (t = 0 s), the active unexposed bacteria deform the interphase boundary and penetrate the
passive region. Over time, active bacteria convect immobile bacteria away from the passive domain, causing
the boundary to erode and propagate inward. The swarm dissolves the passive phase in 60 s, with interface
speeds greater than that of expanding colony edge. (b) The swarm edge (close-up) features densely packed
cells with local polarity and nematic order. (c) The swarm's collective motion recovers after dissolution as
shown by the probability distribution of bacterial speeds p(v), before and after exposure. (d) A montage of
the flow streamlines - from the highlighted box in (a) - reveals the motion of vortices (labeled by color) at
the interface (blue line). Vortices starting in the bulk can collide and attach to the interface (labeled brown
vortex for example); some vortices at the surface detach and move away (green, orange). Others fade away
(purple) or split (dark blue splits from light blue vortex in right tile).
19
FIG. 2. (Color Online) Growth and structure of the active-passive interface in the bacterial
swarm Active-passive domain boundaries are designed with different initial shapes by varying the geometry
of the aperture, such as an octagon [O] (a) or half-space [H] (b). (a-i & b-i) The interface position changes
shape as it moves over time. (a-ii & b-ii) The interface position is identified through order parameters
φ based on fluctuations in image intensity (∆t = 0.1 s) and φv, the bacterial speed (Methods 4). The
interface positions in (ai) and (bi) correspond to φ∗ = 0. For the [O] aperture, the interface roughens as the
domain shrinks in size; for [H], the roughness appears to stabilize. (a-iii & b-iii) We find that averaged
one-dimensional profiles of the φ-fields possess a smooth transitions between the active (φ = 1) and passive
(φ = −1) phases. Fits of the data (dashed lines, Eqn. 6 & 7 Methods) yield the mean location d and width
w of the interface. (a-iv and b-iv) Parameters d and w are determined for multiple experiments (N = 4 per
condition) and averaged together as a function of time (min to max variation in grey). For the [O] aperture,
the mean interface position d(t) (black line, measured radially from the center of the domain) decreases over
t (blue line) at t ≈ 2 sec. Width
w (a-v & b-v) ranges from 4 and 10 µm and varies little over time - except for the case of the octagon
aperture at long times (t > 40 s) as the passive domain dissolves entirely. Scale bar, 50 µm.
time and follows the scaling law d ∼(cid:112)(t0 − t) with t0 ≈ 40 s. For the [H] aperture, the interface position
d initially follows d ∼ t (red line) and then transitions to faster than d ∼ √
20
FIG. 3. (Color Online) Interface structure is coupled to collective flow of bacteria (a) A snap-
shot of the bacterial vorticity field, t = 4 s after exposure with the half-space (H) aperture. Right inset, the
strength of the vorticity field, characterized by (cid:104)ω2(cid:105), increases as one moves away from the interface bound-
ary; while the flow decay timescale τ decreases (grey variation and error bars from temporal fluctuations).
(b) The overall decay of the interface static structure factor scales as ∆hq2 ∼ q−2, yielding an interfacial
stiffness κ = 0.7 µm−2 (N = 4; grey representing min to max variation). (c) The energy spectrum of the
flow E(q) (Equation 9) is non-monotonic with a peak that shifts to higher values of q close to the interface
(Y ≈ 40 µm) compared with the bulk active phase (Y ≈ 130 µm). The peaks in E(q) coincides with peaks
q.
in ∆h2
21
FIG. 4. (Color Online) Evolution of the active-passive interface follows an active analogue to
classical Gibbs-Thomson-Stefan conditions (a) Particle motion at the interface. (i) The trajectories of
2 µm polystyrene tracer particles, reveal how the swarm extracts particles from the passive domain. First,
particles in the passive phase caged-in by their neighbors do not move. Second, when the diffuse interphase
is roughly a velocity correlation length O(wv) away, the trapped particle starts to fluctuate, as seen in the
time trace of its speed (b). Third, agitations induced by the active particles dislodge the jammed tracer from
the cage of neighboring passive particles. For the tracer in (iii), the flow field moves the particle tangentially
along the interface (SI Movie 6). Finally, the particle escapes the passive domain as normal streaming flows
between adjacent vortices pull the particle away, primarily perpendicularly from the interface. This escape
correlates with a rapid increase in particle speed; once inside the swarm, the speed of the tracer fluctuates
(red arrow in ii and iii). The schematic in (iii) defines the sign of the curvature and normal bacterial velocity.
(b,c) Local interface velocity vint correlates with the interface curvature C and the normal component of
the collective bacterial velocity vN at the interface. Data for vint, C, and (vN), gathered over time, collapses
unto a plane for both (b) half-plane [H] and the (c) octagonal [O] apertures. This suggests an active analog
of the classical Gibbs-Thomson- Stefan boundary conditions exists.
|
1512.00496 | 3 | 1512 | 2016-03-09T02:59:33 | Sense and sensitivity: physical limits to multicellular sensing, migration and drug response | [
"physics.bio-ph",
"q-bio.CB"
] | Metastasis is a process of cell migration that can be collective and guided by chemical cues. Viewing metastasis in this way, as a physical phenomenon, allows one to draw upon insights from other studies of collective sensing and migration in cell biology. Here we review recent progress in the study of cell sensing and migration as collective phenomena, including in the context of metastatic cells. We describe simple physical models that yield the limits to the precision of cell sensing, and we review experimental evidence that cells operate near these limits. Models of collective migration are surveyed in order understand how collective metastatic invasion can occur. We conclude by contrasting cells' sensory abilities with their sensitivity to drugs, and suggesting potential alternatives to cell-death-based cancer therapies. | physics.bio-ph | physics |
Sense and sensitivity: physical limits to multicellular sensing, migration and drug
response
Julien Varennes and Andrew Mugler∗
Department of Physics and Astronomy, Purdue University, West Lafayette, IN 47907, USA
Metastasis is a process of cell migration that can be collective and guided by chemical cues. View-
ing metastasis in this way, as a physical phenomenon, allows one to draw upon insights from other
studies of collective sensing and migration in cell biology. Here we review recent progress in the
study of cell sensing and migration as collective phenomena, including in the context of metastatic
cells. We describe simple physical models that yield the limits to the precision of cell sensing, and
we review experimental evidence that cells operate near these limits. Models of collective migration
are surveyed in order understand how collective metastatic invasion can occur. We conclude by con-
trasting cells' sensory abilities with their sensitivity to drugs, and suggesting potential alternatives
to cell-death-based cancer therapies.
Keywords: metastasis, collective behavior, gradient sensing, cell migration
Metastasis is one of the most intensely studied stages
of cancer progression because it is the most deadly stage
of cancer. The first step of metastasis is invasion, wherein
cells break away from the tumor and invade the surround-
ing tissue. Our understanding of metastatic invasion has
benefited tremendously from genetic and biochemical ap-
proaches [25, 26, 36]. However, the physical aspects of
metastatic invasion are still unclear [26]. We know that
at a fundamental level, metastatic invasion is a physi-
cal process. Tumor cells sense and respond to chemical
gradients provided by surrounding cells [4, 14, 51, 55]
or other features of the tumor environment [49, 54, 55]
(Fig. 1A). Indeed, tumor cells are highly sensitive, able
to detect a 1% difference in concentration across the cell
length [55]. Sensing is ultimately a physical phenomenon.
Therefore, can we build a simple physical theory to un-
derstand the sensory behavior of tumor cells, and can
this physical theory inform treatment options?
Metastatic invasion involves coordinated migration of
tumor cells away from the tumor site. In many types of
cancer, migration is collective and highly organized, in-
volving the coherent motion of connected groups of cells
[1, 11, 21, 51] (Fig. 1B). Collective migration is ultimately
a physical phenomenon, since it relies on mechanical cou-
pling and can often be understood as emerging from sim-
ple physical interactions at the cell-to-cell level. Can we
understand the collective migration of tumor cells with
simple physical models?
Here we review recent progress on modeling sensing
and migration in cells and cell collectives. We discuss
metastatic cells explicitly, and emphasize that physical
insights gained from other cellular systems can inform
our understanding of metastatic invasion. We focus on
simple physical models and order-of-magnitude numeri-
cal estimates in order to quantitatively probe the extent
of, and the limits to, cell sensory and migratory behavior.
Our hope is that a more quantitative understanding of
∗Electronic address: [email protected]
Figure 1: Metastatic invasion is guided by chemical attrac-
tants and can occur via (A) single cells or (B) multicellular
groups. (C) Drugs are delivered to the tumor environment in
order to prevent tumor growth and metastasis. Drugs may
cause cell death (orange), block cell-to-cell communication
(purple), or prevent cell migration (blue).
metastatic invasion will inform treatment protocols, and
to that end we conclude by discussing drug sensitivity
and potential treatment strategies (Fig. 1C).
I. PHYSICAL LIMITS TO SENSORY
PRECISION
cells
sense
very
small
Tumor
concentration
gradients[55] and act in a collective manner[1, 11, 21, 51].
Here we review the basic theory of concentration and
gradient sensing by cells and cell collectives. This theory
places physical bounds on sensory precision and allows
us to quantitatively compare the capabilities of tumor
cells to other cell types.
A. Single-cell concentration sensing
Theoretical
limits to the precision of concentration
sensing were first introduced by Berg and Purcell almost
40 years ago[3]. Berg and Purcell began by considering an
idealized cell that acts as a perfect counting instrument.
Their simplest model assumed that the cell is a sphere
in which molecules can freely diffuse in and out (Fig.
2A). The concentration of these molecules is uniform in
space, and the cell derives all its information about the
concentration by counting each molecule inside its spher-
ical body. The expected count is ¯n = ¯cV where ¯c is the
mean concentration and V is the cell volume. However,
since molecules arrive and leave via diffusion, there will
be fluctuations around this expected value. Diffusion is a
Poisson process, meaning that the variance in this count
√
σ2
n equals the mean ¯n. Therefore the relative error in the
cell's concentration estimate is σc/¯c = σn/¯n = 1/
¯cV .
The cell can improve upon the relative error in its con-
centration estimate by time-averaging over multiple mea-
surements. However, consecutive measurements are only
statistically independent if they are separated by a suffi-
cient amount of time such that the molecules inside the
cell volume are refreshed. The amount of time required is
characterized by the diffusion time, τ ∼ V 2/3/D ∼ a2/D,
where D is the diffusion constant and a is the cell diame-
ter. In a time period T the cell makes ν = T /τ indepen-
dent measurements, and the variance is reduced by the
factor 1/ν. This gives the long-standing lower limit
σc
¯c
=
∼
σn
¯n
1√
a¯cDT
(1)
for the cell's relative error in estimating a uniform con-
centration. The relative error decreases with a and ¯c,
since the molecule count is larger, and also with D and
T , since more independent measurements can be made.
Berg and Purcell derived this limit more rigorously[3],
and the problem has been revisited more recently to ac-
count for binding kinetics, spatiotemporal correlations,
and spatial confinement [5, 6, 31]. In all cases a term of
the form in Eq. 1 emerges as the fundamental limit for
three-dimensional diffusion.
Do cells reach this limit? Berg and Purcell them-
selves asked this question in the context of several single-
celled organisms, including the Escheria coli bacterium
[3]. Motility of E. coli has two distinct phases: the run
phase in which a cell swims in a fixed direction, and the
tumble phase in which the cell erratically rotates in or-
der to begin a new run in a different direction. The bac-
terium biases its motion by continually measuring the
chemoattractant concentration, and extending the time
of runs for which the change in concentration is positive
[3, 9, 15, 61]. The change in concentration ∆¯c = T v¯g
over a run time T depends on the concentration gradi-
ent ¯g = ∂¯c/∂x and the bacterium's velocity v. Berg and
2
Purcell argued that for a change in concentration to be
detectable, it must be larger than the measurement un-
certainty, ∆¯c > σc. Together with Eq. 1, this places a
lower limit on the run time, T > [¯c/(aDv2¯g2)]1/3. Us-
ing typical values [3] for the sensory threshold of E. coli
of ¯c = 1 mM, ∂¯c/∂x = 1 mM/cm, a = 1 µm, v = 15
µm/s, and D = 10−5 cm2/s, we find T > 0.1 s. Ac-
tual run times are on the order of 1 s. Thus we see
that E. coli chemotaxis is consistent with this physical
bound. Although the end goal of concentration sensing
in E. coli is chemotaxis by temporally sampling changes
in the chemical concentration, we would like to focus the
reader's attention on the remarkable fact that the bac-
terium's concentration sensing machinery operates very
near the predicted physical limits. If E. coli were to use
any shorter run times, chemotaxis would be physically
impossible. Consequently, the time period for measuring
the chemical concentration, T in Eq. 1, would be so short
that the bacterium would be unable to make an accurate
measurement of the chemical concentration.
B. Single-cell gradient sensing
Cells are not only able to detect chemical concentra-
tions, they are also able to measure spatial concentration
gradients. Many cells, including amoeba, epithelial cells,
neutrophils, and neurons, sense gradients by comparing
concentrations between compartments in different loca-
tions of the cell body [29]. These compartments are typ-
ically receptors or groups of receptors on the cell surface,
but in a simple model we may treat these compartments
as idealized counting volumes as we did before (Fig. 2B).
The difference in counts between two such compartments
provides the cell with an estimate of the gradient. What
is the relative error in this estimate?
Consider two compartments of linear size s on either
side of a cell with diameter a (Fig. 2B). If the compart-
ments are aligned with the gradient ¯g of a linear con-
centration profile, then the mean concentrations at each
compartment are ¯c1 and ¯c2 = ¯c1+a¯g. The mean molecule
counts in the two compartments are roughly ¯n1 = ¯c1s3
and ¯n2 = ¯c2s3, and the difference is ∆¯n = ¯n2−¯n1 = a¯gs3.
∼
The variance in this difference is σ2
1/(s¯c1DT ) + ¯n2
¯n2
2/(s¯c2DT ), where the first step assumes
the two compartments are independent, and the sec-
ond step uses Eq. 1 for the variance in each compart-
ment's measurement. For shallow gradients, where the
limits on sensing are generally probed, we have a¯g (cid:28) ¯c1,
and therefore we may assume ¯c1 ≈ ¯c2 ≈ ¯c, where ¯c is
the mean concentration at the center of the cell. Thus
∆n ∼ 2(¯cs3)2/(s¯cDT ), and the relative error in the cell's
σ2
estimate of the gradient is then
∆n = σ2
n1
+ σ2
n2
(cid:114)
σg
¯g
=
∼
σ∆n
∆¯n
¯c
s(a¯g)2DT
,
(2)
where the factor of 2 is neglected in this simple scaling
estimate. As in Eq. 1, we see that the relative error
3
Figure 2:
(A) An idealized cell as a permeable sphere that counts molecules inside its volume. (B) A cell counts molecules
in two compartments in order to estimate a concentration gradient. (C) The local excitation–global inhibition (LEGI) model
of multicellular gradient sensing. Y molecules diffuse between neighboring cells, whereas X molecules do not. The difference
between X and Y counts in a given cell reports the extent to which that cell's concentration measurements are above the
average.
decreases with s, since the molecule counts in each com-
partment are larger, and also with D and T , since more
independent measurements can be made. Additionally,
the relative error decreases with a¯g, since the concentra-
tions measured by the two compartments are more differ-
ent from each other. However, we see that unlike in Eq. 1,
the relative error increases with the background concen-
tration ¯c. The reason is that the cell is not measuring a
concentration, but rather a difference in concentrations,
and it is more difficult to measure a small difference on
a larger background than on a smaller background [17].
Eq. 2 has been derived more rigorously[19], and the prob-
lem has been extended to describe rings of receptors [19]
or detectors distributed over the surface of a circle [27]
or a sphere [18]. In all cases a term of the form in Eq. 2
emerges as the fundamental limit, with the lengthscale s
dictated by the particular sensory mechanism and geom-
etry. It is clear that the optimal mechanism would result
in an effective compartment size that is roughly half of
the cell volume, in which case s ∼ a.
Do cells reach this limit on gradient sensing? This
question has been directly addressed for the amoeba Dic-
tyostelium discoideum. Experiments[64] have shown that
Dictyostelium cells exhibit biased movement when ex-
posed to gradients of cyclic adenosine monophosphate as
small as ¯g = 10 nM/mm, on top of a background concen-
tration of ¯c = 7 nM. Bias is typically quantified in terms
of the chemotactic index (CI), which is the cosine of the
angle between the gradient direction and the direction of
a cell's actual motion. By relating the error in gradient
sensing (a term of the form in Eq. 2 with s = a) to the
error in this angle, Endres and Wingreen [18] obtained
an expression for the optimal CI, which they then fit to
the experimental data with one free parameter, the inte-
gration time T . The inferred value of T = 3.2 s serves as
the physical lower bound on the response time required
to perform chemotaxis. Actual response times of Dic-
tyostelium cells, as measured by the time from the addi-
tion of a chemoattractant to the peak activity of an ob-
servable signaling pathway associated with cell motility
[48, 50], are about 5−10 s. Taken together, these results
imply that Dictyostelium operates remarkably close to
the physical limit to sensory precision set by the physics
of molecule counting.
C. Relative changes vs. absolute molecule numbers
The precision of gradient sensing is often reported in
terms of percent concentration change across a cell body.
For example, both amoeba [64] and tumor cells [55] are
sensitive to a roughly 1% change in concentration across
the cell body. However, this method of reporting sensi-
tivity may be misleading. Experiments imply very differ-
ent sensory thresholds for these cells in terms of absolute
molecule numbers, as we will now see.
The key is that it takes two numbers to specify the
conditions for gradient sensing: the mean gradient ¯g and
the mean background concentration ¯c. For the amoeba
Dictyostelium, these numbers are ¯g = 10 nM/mm and
¯c = 7 nM at the sensory threshold [64]. Given a typical
cell size of a = 10 µm, these values imply a mean percent
concentration change of ¯p = a¯g/¯c = 1.4% (Table I). How-
ever, we may also compute from these values the mean
molecule number difference ∆¯n = a¯gs3 from one side of
the cell to the other, within the effective compartments
of size s. Taking s ∼ a gives the maximal molecule num-
ber difference of ∆¯n = a4¯g = 60 for Dictyostelium (Table
I). Together ¯p and ∆¯n specify the sensing conditions as
completely as ¯g and ¯c do.
Experiments [55] have shown that breast cancer tumor
cells exhibit a chemotactic response in a gradient ¯g = 550
nM/mm of the cytokine CCL21, on top of a background
concentration of ¯c = 1100 nM. Given a typical cell size
of a = 20 µm, this corresponds to a percent difference of
¯p = a¯g/¯c = 1%, similar to Dictyostelium. Yet, this also
corresponds to a maximal molecule number difference of
∆¯n = a4¯g = 53,000, which is much higher than that of
Dictyostelium (Table I). Even though the sensitivities are
similar in terms of percent change, they are very different
in terms of absolute molecule number.
Lower molecule numbers correspond to higher relative
error. We can see this explicitly by writing Eq. 2 in
terms of the percent change ¯p = a¯g/¯c. Defining = σg/¯g
and taking s ∼ a, we have ∼ 1/(cid:112)¯p2a¯cDT . Ac-
counting for the fact that tumor cells (TC) have roughly
twice the diameter as Dictyostelium cells (DC), this ex-
pression implies that the sensitivities of the two cell
types over the same integration time T to chemoat-
tractants with the same diffusion constant D satisfy
DC/TC =(cid:112)2¯cTC/¯cDC ≈ 18. We see that because the
Dictyostelium experiments were performed at lower back-
ground concentration, corresponding to lower absolute
molecule numbers, the relative error in gradient sensing
is 18 times that of the tumor cells, despite the fact that
both cell types are responsive to 1% concentration gradi-
ents. Therefore, it is important to take note of the back-
ground concentration when studying the precision of gra-
dient sensing. These data imply that Dictyostelium cells
can sense noisier gradients than tumor cells. However,
Dictyostelium cells have been studied more extensively
than tumor cells as exemplars of gradient detection. It
remains an interesting open question what is the mini-
mum gradient that tumor cells can detect, not only in
terms of percent concentration change, but also in terms
of absolute molecule number differences.
We see that although cancerous cells and Dictyostelium
cells are of similar size, their sensory responses to abso-
4
lute molecule numbers can be very different (Table I).
This difference is also reflected in their migration speeds:
carcinoma and epithelial cells migrate [22, 37, 65, 66] at
∼ 0.5µm/s whereas Dictyostelium can migrate [45, 57] at
speeds of ∼ 10µm/s.
D. Multicellular gradient sensing
In many cancer types, tumor cells invade the surround-
ing tissue in a collective manner [11, 21]. Cell collectives
can sense shallower gradients than single cells [17, 42],
both in terms of percent concentration changes and ab-
solute molecule numbers (Table I). Indeed, groups of neu-
rons respond to gradients equivalent to a difference of less
than one molecule across an individual neuron's growth
cone [52]. This raises the possibility that during the in-
vasion process tumor cell collectives benefit from higher
sensory precision than single tumor cells.
We can understand immediately from Eq. 2 why a
multicellular collective would have lower sensory error
than a single cell: a collective is larger than a single cell.
Therefore, the collective covers a larger portion of the
concentration profile, which leads to a larger difference
between the concentration measurements on either end,
and a lower relative error. In terms of Eq. 2, if we con-
sider that cells on the ends act as the molecule-counting
compartments, s → a, and that the entire collective acts
as the detector, a → N a, where N is the number of cells
in the gradient direction, then we have [47]
(cid:114)
∼
σg
¯g
¯c
a(N a¯g)2DT
.
(3)
We see that, as expected, the relative error goes down
with the size N a of the multicellular collective.
However, there is a crucial point that is overlooked
in formulating Eq. 3: the larger the group of cells, the
more difficult it is for cells on either end to communicate
the measurement information. This fact is not accounted
for in Eq. 3. Instead, we see that the relative error de-
creases with the separation N a between the end cells
without bound, which is unrealistic. For a single cell it
may be a reasonable approximation to assume that com-
partments quickly and reliably communicate information
across the cell body, but for a multicellular collective,
the communication process cannot be overlooked.
Im-
portantly, the communication mechanism of multicellular
collectives may introduce additional noise into the gra-
dient sensing process. Therefore, it is imperative when
considering collective sensing to properly account for the
effects of communication.
Recently, the physical limits to collective gradient sens-
ing including communication effects were derived [17, 47].
Communication was modeled using a multicellular ver-
sion of the local excitation–global
inhibition (LEGI)
paradigm [38], in which each cell produces a "local" and
a "global" molecular species in response to the chemoat-
tractant, and the global species is exchanged between
5
Single Cell
Multicellular
Dictyostelium
(Amoeba) [64] Cancer [55]
Breast
Neurons [52] Mammary
Epithelia [17]
10 µm
20 µm
10 µm
10 µm
7 nM
1100 nM
1 nM
2.5 nM
10 nM/mm 550 nM/mm 0.1 nM/mm 0.5 nM/mm
1.4%
1.0%
0.1%
0.2%
60
53,000
0.6
3
Cell Length
Scale, a
Background
Concentration, ¯c
Concentration
Gradient, ¯g
Percent Concentration
Difference, ¯p = ¯ga/¯c
Molecule Number
Difference, ∆¯n = ¯ga4
Table I: Gradient sensory thresholds for single cells and multicellular collectives. Note that experiments can provide equal
percent concentration differences but unequal molecule number differences across a cell body, as seen for amoeba and breast
cancer cells. We see that multicellular groups can detect smaller gradients than single cells by all measures.
cells to provide the communication (Fig. 2C). The dif-
ference between local and global molecule numbers in a
given cell provides the readout. A positive difference in-
forms the cell that its detected chemoattractant concen-
tration is above the spatial average among its neighbors,
and therefore that the cell is located up the gradient,
not down. In this model, the relative error of gradient
sensing was shown [47] to be limited from below by
(cid:114)
∼
σg
¯g
¯c
a(n0a¯g)2DT
,
(4)
where n2
0 is the ratio of the cell-to-cell exchange rate to
the degradation rate of the global species. Comparing
Eq. 3 to Eq. 4, we see that without communication the
error decreases indefinitely with the size N a of the collec-
tive, whereas with communication the error is bounded
by that of a collective with effective size n0a. Evidently,
communication defines an effective number of cells n0
over which information can be reliably conveyed, and a
collective that grows beyond this size no longer improves
its sensory precision.
These theoretical predictions were tested experimen-
tally in collectives of epithelial cells [17]. Mouse mam-
mary epithelial cells were grown in organotypic culture
and subjected to very shallow gradients of epidermal
growth factor (Table I). It was shown that while single
cells did not respond to these gradients, the multicellu-
lar collectives did: they exhibited a biased cell-branching
response. Importantly, the response of large collectives
was no more biased than the response of small collec-
tives, supporting the theory with communication (Eq. 4)
over the theory without communication (Eq. 3). The ef-
fective detector size was inferred to be n0 ≈ 3.5 cells,
which is consistent with the size of these collectives in
their natural context (the "end buds" of growing mam-
mary ducts) [39]. Interestingly, when the gap junctions
between cells, which mediate the molecular communica-
tion, were blocked with each of several drugs, the bi-
ased responses were abolished [17], demonstrating that
the collective response was critically dependent on the
cell-to-cell communication. Taken together, these results
indicate that cell-to-cell communication is a necessary
but imperfect enabler of collective gradient sensing. The
results also speak to the power of simple physical theory
to quantitatively explain collective cellular capabilities.
Since epithelial cancers are known to invade collectively
[11], it remains an important open question whether this
theory also describes the sensory behavior of tumor cell
collectives.
II. PHYSICAL MODELS OF COLLECTIVE
MIGRATION
Metastatic invasion is a process of cell migration. Col-
lective invasion, in turn, is a process of collective migra-
tion. Therefore, it is important to understand not only
the collective sensing capabilities of tumor cells, but also
the properties of their collective migration-and ideally
the relation between the two. From a physical modeling
perspective, describing collective cell dynamics is an in-
teresting problem, because often rich and unexpected be-
havior can emerge from a few simple interaction rules be-
tween cells. Even in the absence of sensing, simple models
have successfully explained observed collective behaviors
such as cell streaming, cell sorting, cell sheet migration,
wound healing, and cell aggregation [2, 28, 30, 59]. Here
we focus on the collective dynamics that emerge when
sensing plays a key role. In this case, a sensory cue re-
sults in polarization of a cell or cell collective via one
of a variety of mechanisms [29], and the dynamics are
directed, i.e. migratory.
A. Mechanisms of collective migration
Broadly speaking, the mechanisms of collective migra-
tion can be divided into three categories. First, cells may
exhibit individual sensing and individual migration (Fig.
3A). Here, each cell can perform gradient sensing and mi-
gration individually, although the precision may be low.
When many such cells are placed in a group, the group
migration can be enhanced and focused by local interac-
tions between the cells. Even if each individual cell has
low sensory and migratory precision, the precision of the
group as a whole is high due to the interactions. Colli-
sions act to average over the errors in individual cells'
noisy measurements, thereby decoupling group behav-
ior from single-cell properties. This mechanism is often
termed "many wrongs," and it is successful at explaining
how group migratory behavior emerges from individual
agents that act independently [13, 56]. As discussed later,
the failure of a communication-blocking therapy could
act as proof that a "many wrongs" method of collective
migration is at work in tumor cell invasion.
Second, cells may exhibit individual sensing but col-
lective migration (Fig. 3B). In this mechanism, each in-
dividual cell senses its own local environment, and tight
mechanical interactions result in the emergent directed
motion of the entire group. This mechanism is applicable
to the collective migration of connected clusters of cells.
For example, models of this type were recently devel-
oped by Camley et al. [8] and by Malet-Engra et al. [42]
to describe behavior seen in clusters of neural crest cells
and lymphocytes, respectively. In this model, cells are
tightly connected but are polarized away from neighbor-
ing cells due to contact inhibition of locomotion (CIL),
the physical phenomenon of cells ceasing motion in the
direction of cell-cell contact [44]. Individual cells sense a
local chemoattractant concentration and attempt to mi-
grate away from the group with a strength proportional
to this concentration. However, the mechanical coupling
keeps them together.
In the presence of a concentra-
tion gradient, the imbalance in their migration strengths
results in net directed motion (Fig. 3B). Notably, this
mechanism results in directed motion of a cluster even
though individual cells cannot execute directed motion
alone, since without other cells, there is no CIL to bias
the motion.
Third, cells may exhibit collective sensing and collec-
tive migration (Fig. 3C). As discussed above, multicel-
lular groups exploit cell-to-cell communication to sense
gradients collectively, thereby enhancing the precision
of sensing. A feature of this collective sensing, e.g.
via the multicellular LEGI mechanism discussed above
[17, 47], is that each cell has information on the extent
to which it is up or down the gradient. Through CIL
or other contact-mediated interactions, this information
can translate directly into cell polarity, leading to more
coherent collective migration than in the previous mech-
anism (Fig. 3C vs. B). In fact, the multicellular LEGI
model was used by Camley et al. [8] to explore a model
6
of this type. Adding collective sensing to their model
of CIL-dependent migration gave the advantage that the
repulsive tension on a cell cluster was adaptive and there-
fore remained constant as the cluster migrated to regions
of higher chemical concentration.
B. Model implementations
To study the above mechanisms quantitatively and
compare predictions with experiments, one must turn to
mathematical and computational modeling. Models of
cell dynamics range from continuum or semi-continuum
descriptions, which describe groups of cells as continu-
ous tissues, to individual-based models, which describe
cells as individual
interacting entities [40]. Physics-
driven individual-based models generally fall into two
categories: force-based models and energy-based models.
Force-based models (Fig. 3D) typically represent cells
as centers of mass or as collections of vertices. Cell
dynamics evolve from forces acting on individual cells,
which can be stochastic, and arise from internal fea-
tures such as cell polarity, and external features such
as mechanical interactions with other cells [40]. Force-
based models are able to reproduce multicellular behavior
such as chemotaxis, wound healing, and cell aggregation
[2, 8, 28]. Parameters are often directly relatable to ex-
perimental measurements, and the simplest models are
often amenable to exact mathematical analysis [8].
Energy-based models (Fig. 3E) allow cell dynamics to
emerge from the minimization of a potential energy with
thermal noise (the so-called Monte Carlo scheme). A
widely used example is the cellular Potts model (CPM)
[23, 58], in which cells are represented as collections of
co-aligned "spins" on a lattice (Fig. 3E). Cells remain
contiguous because it is energetically favorable for neigh-
boring spins to be co-aligned. Biophysical features such
as cell shape, cell-cell adhesion, and cell protrusions into
the environment are modeled by introducing correspond-
ing terms into the global potential energy. The CPM
has successfully reproduced experimental observations of
cell sorting, streaming, chemotaxis, and collective migra-
tion [30, 43, 59]. In energy-based models, the parameters
are set by calibrating emergent features, such as cell dif-
fusion coefficients or average speeds, with experimental
measurements [59].
Although the physical limits to multicellular sensing
are becoming better understood, the physical limits con-
straining multicellular migration are less clear. This re-
mains an interesting open question, and answering it will
require integrating the theories of sensing and communi-
cation with the models of collective migration described
herein. For tumor cells in particular, an integrated phys-
ical theory of sensing and migration would prove im-
mensely useful for identifying the key determinants of in-
vasive capabilities. Identifying these determinants would
help pinpoint the ways that these capabilities could be
disrupted, using drugs and other therapies, as described
7
Figure 3: Mechanisms of collective migration: (A) individual sensing and migration (the "many wrongs" mechanism), (B)
individual sensing but collective migration (emergent chemotaxis), and (C) collective sensing and migration. Implementations
of collective migration: (D) in force-based models, dynamics evolve from stochastic forces acting on each cell; (E) in energy-
based models, dynamics evolve via energy minimization with thermal noise. E shows the cellular Potts model framework, in
which cells are collections of lattice sites, and cell-cell (dashed blue) and cell-environment (dashed yellow) contacts contribute
to the energy of the system.
next.
III. DRUG SENSITIVITY AND IMPLICATIONS
FOR THERAPY
We have seen that cells, including tumor cells, are re-
markably precise sensors of molecules in their environ-
ment. This raises the question of how sensitive tumor
cells are to drug molecules in their environment. What
is the minimum drug concentration required not just for
precise detection by a cell, but for causing a phenotypic
change, such as cell death?
Experiments have shown that cancer cells are sensitive
to very small drug concentrations. For example, lung
carcinoma cells were exposed in vitro to various concen-
trations of the anti-cancer drug paclitaxel, also known as
taxol, which acts to block mitosis in order to achieve cell
death by disrupting microtubule regulation [60]. Pacli-
taxel concentrations as low as 1 nM were shown to affect
microtubule dynamics of the cells. This concentration is
commensurate with the smallest background concentra-
tions in which cells can perform gradient sensing (Table
I). Assuming a cell length of 20 µm, which is typical of
carcinoma cells [36], this concentration corresponds to
only a few thousand drug molecules in the volume of a
cell (Table II). Evidently lung cancer cells are affected by
drug concentrations that are near the fundamental limits
of what can be sensed.
Although cancer cells may be very sensitive to small
drug concentrations, that does not translate to successful
treatment.
In order to achieve cell death, much larger
drug concentrations are required. In the same study on
lung carcinoma cells, cell death was observed for drug
concentrations on the order of 10 nM and greater. More
typical drug concentrations required for cell death are
on the order of micromolars. For instance, it has been
shown in vitro that anticancer drug concentrations on
the order of 10 µM are required to kill at least 90% of
Outcome
Drug Concentration Molecules per Cell
Physical change [60]
Cell death [24]
Cell death, nanoparticle delivery [41]
Communication blockage [17]
1 nM
104 nM
100 nM
50 nM
5,000
5 × 107
500,000
30,000
8
Table II: Drug sensitivity thresholds. Molecules per cell volume are calculated assuming a cubic cell of length a = 20 µm for
tumor cells (rows 1-3) and a = 10 µm for epithelial cells (row 4).
tumor cells [24]. With a cell length of 20 µm, 10 µM
corresponds to tens of millions of drug molecules in the
volume of a cell, four orders of magnitude greater than
drug concentrations required to affect cell functionality
(Table II). In order to effectively kill a solid tumor, very
high drug doses are required.
Complicating matters is the fact that the tumor and its
surrounding microenvironment comprise a complex and
heterogeneous system. Although most cells in the human
body are naturally within a few cell lengths of a blood
vessel, due to high proliferation tumor cells may be up-
wards of tens of cell lengths away from a vessel [46]. This
makes it difficult for drugs to reach the tumor. Moreover,
the high density of many solid tumors causes gradients
of drug concentration to form as a function of tumor ra-
dius [34]. This results in a reduced drug concentration at
the center of the tumor and makes innermost tumor cells
the most difficult to kill. A promising way to overcome
this difficulty is through the use of nanoparticle drug de-
livery systems, which increase both the specificity and
penetration of drugs to the tumor. Nanoparticle delivery
has been shown [41] to achieve cell death with concentra-
tions as low as 100 nM. Although this concentration is
lower than delivery without nanoparticles, it is still two
orders of magnitude higher than the minimum concen-
tration that causes physical changes in the cell (Table
II). Even with targeted delivery, achieving drug-induced
tumor cell death remains a challenging task.
Given this challenge, we hope to draw upon the physi-
cal insights reviewed herein to devise therapeutic strate-
gies that are alternative or complementary to comprehen-
sive cell death. Specifically, we imagine focusing on the
metastatic invasion phase, and targeting the functions of
invading tumor cells, including communication and mi-
gration, in addition to targeting cells' overall viability, to
produce better treatment (Fig. 1C). Communication is
a particularly promising candidate, since it has recently
been shown that cell-to-cell communication makes cancer
cells more resistant to therapy and helps sustain tumor
growth [7]. Indeed, the exchange of extracellular vesicles,
which is a form of communication observed between tu-
mor cells and stromal cells, has been linked to immune
suppression, tumor development, angiogensis, and metas-
tasis [62]. This suggests that disrupting cell-to-cell com-
munication could be an effective strategy for stopping
tumor progression or curbing metastatic invasion. Dis-
rupting communication may not require concentrations
as large as those necessary for cell death, which are dif-
ficult to maintain in vivo across the whole tumor. For
example, as little as 50 nM of the gap-junction-blocking
drug Endothelin-1 is sufficient to remove collective re-
sponses in epithelial cells [17]. This concentration is sev-
eral orders of magnitude smaller than that required for
comprehensive cell death, and it is on the order of con-
centrations that are effective with targeted nanoparticle
delivery (Table II). Therefore, it is tempting to suggest
that managing metastatic invasion by blocking commu-
nication or other cell functions is a more accessible ther-
apeutic strategy than eradicating a tumor outright.
Although blocking intercellular communication path-
ways could curb the invasive potential of metastatic cells
it is also important to address the ulterior consequences
of this strategy. Gap junction intercellular communica-
tion (GJIC) is an important way for the environment to
affect change on cells, maintaining tissue homeostasis and
balancing cell growth and death [33]. In cancerous cells
GJIC is reduced, causing unregulated cell growth [53].
Interestingly, many existing cancer-combatting drugs are
small enough to pass through cell gap junctions which
permit molecules of sizes up to 1000 Dalton, but there
is a lack of in vivo studies concerning the benefits and
effects of gap junctions on cancer treatment [53]. It has
been shown in vitro that GJIC can propagate cell-death
signals through cancerous cells and that high connexin
expression, the proteins that compose gap junctions, cor-
responds to high anticancer drug sensitivity [32, 33].
Therefore, it is important to consider the potential neg-
ative consequences of blocking intercellular communica-
tion in reducing metastatic invasion. It may be sufficient
to administer an anticancer drug and a communication-
blocking drug at different times in order to avoid negative
side-effects. Although this puts a caveat on our proposal
of communication-blocking drugs as a viable option for
treating metastatic invasion it is important to recall that
GJIC is not the only communication pathway available
to cancerous cells: extracellular vesicle-meditated signal-
ing pathways are potential alternates which could be tar-
geted in place of GJIC [7, 62].
IV. OUTLOOK
In this review, we have taken a quantitative look at
metastatic invasion as a sensing-and-migration process,
which has allowed us to compare metastatic cells to other
cell types in terms of their physical capabilities. We have
seen that tumor cells can sense very shallow chemoattrac-
tant gradients, which may help guide metastatic invasion,
but it remains unclear whether tumor cells operate near
fundamental sensing limits, as bacteria and amoeba do.
Recognizing that metastatic invasion can be collective,
we have reviewed recent results on the physical limits to
collective sensing, and we have identified the overarching
mechanisms of collective migration. A key insight that
emerges is that collective capabilities rely critically on
cell-to-cell communication. This insight opens up alter-
native strategies for therapy that target specific cell capa-
bilities such as communication, in addition to strategies
that aim for comprehensive cell death.
A detailed presentation of the underlying physical me-
chanics for cell motility and chemotaxis are outside the
scope of this review. Readers interested in these topics
are referred to these excellent resources [12, 20, 35, 63].
It is also important to note that in deriving the lim-
its to concentration sensing we have assumed that the
molecules of interest diffuse normally with fixed, space-
independent diffusion coefficients. However, this may
not always be the case in the tumor environment,
where molecules can also experience anomalous diffusion
9
[10, 16].
Moving forward,
it will be important to identify
whether the physical theory of sensing reviewed herein
can be applied in a predictive manner to tumor cells, and
whether gradient sensing plays a dominant role during
metastatic invasion. More generally, it will be necessary
to integrate the theory of sensing with models of collec-
tive migration to predict quantitatively what groups of
migratory cells can and cannot do. Finally, controlled ex-
periments with metastatic cells are required to validate
these predictions, and to assess the viability of alterna-
tive therapies that target specific cell functions in order
to combat metastatic invasion. Our hope is that the in-
tegrated, physics-based perspective presented herein will
help generate innovative solutions to the pervasive prob-
lem of metastatic disease.
V. ACKNOWLEDGMENTS
This work was supported by the Ralph W. and Grace
M. Showalter Research Trust and the Purdue Research
Foundation.
[1] Nicola Aceto, Aditya Bardia, David T Miyamoto,
Maria C Donaldson, Ben S Wittner, Joel A Spencer, Min
Yu, Adam Pely, Amanda Engstrom, Huili Zhu, Brian W
Brannigan, Ravi Kapur, Shannon L Stott, Toshi Sh-
ioda, Sridhar Ramaswamy, David T Ting, Charles P Lin,
Mehmet Toner, Daniel A Haber, and Shyamala Mah-
eswaran. Circulating tumor cell clusters are oligoclonal
precursors of breast cancer metastasis. Cell, 158(5):1110–
1122, 2014.
[2] Markus Basan, Jens Elgeti, Edouard Hannezo, Wouter-
Jan Rappel, and Herbert Levine. Alignment of cel-
lular motility forces with tissue flow as a mechanism
for efficient wound healing. Proceedings of the National
Academy of Sciences, 110(7):2452–2459, 2013.
[3] Howard C Berg and Edward M Purcell. Physics of
chemoreception. Biophysical journal, 20(2):193, 1977.
[4] Neil A Bhowmick, Eric G Neilson, and Harold L Moses.
Stromal fibroblasts in cancer initiation and progression.
Nature, 432(7015):332–337, 2004.
[5] William Bialek and Sima Setayeshgar. Physical limits
to biochemical signaling. Proceedings of the National
Academy of Sciences of the United States of America,
102(29):10040–10045, 2005.
[6] Brendan A Bicknell, Peter Dayan, and Geoffrey J Good-
hill. The limits of chemosensation vary across dimensions.
Nature communications, 6, 2015.
[7] Mirjam C Boelens, Tony J Wu, Barzin Y Nabet, Bihui
Xu, Yu Qiu, Taewon Yoon, Diana J Azzam, Christina
Twyman-Saint Victor, Brianne Z Wiemann, Hemant Ish-
waran, Petra J ter Brugge, Jos Jonkers, Joyce Slinger-
land, and Andy J Minn. Exosome transfer from stromal
to breast cancer cells regulates therapy resistance path-
ways. Cell, 159(3):499–513, 2014.
[8] Brian A Camley, Juliane Zimmermann, Herbert Levine,
and Wouter-Jan Rappel. Emergent collective chemo-
taxis without single-cell gradient sensing. arXiv preprint
arXiv:1506.06698, 2015.
[9] Antonio Celani and Massimo Vergassola.
Bacterial
strategies for chemotaxis response. Proceedings of the
National Academy of Sciences, 107(4):1391–1396, 2010.
[10] Vikash P Chauhan, Triantafyllos Stylianopoulos, Yves
Boucher, and Rakesh K Jain. Delivery of molecular and
nanoscale medicine to tumors:
transport barriers and
strategies. Annual review of chemical and biomolecular
engineering, 2:281–298, 2011.
[11] Kevin J Cheung, Edward Gabrielson, Zena Werb, and
Andrew J Ewald. Collective invasion in breast can-
cer requires a conserved basal epithelial program. Cell,
155(7):1639–1651, 2013.
[12] Chang Y Chung, Satoru Funamoto, and Richard A Fir-
Signaling pathways controlling cell polarity and
tel.
chemotaxis. Trends in biochemical sciences, 26(9):557–
566, 2001.
[13] Luke Coburn, Luca Cerone, Colin Torney, Iain D Couzin,
and Zoltan Neufeld. Tactile interactions lead to coher-
ent motion and enhanced chemotaxis of migrating cells.
Physical biology, 10(4):046002, 2013.
[14] John Condeelis and Jeffrey W Pollard. Macrophages:
obligate partners for tumor cell migration, invasion, and
metastasis. Cell, 124(2):263–266, 2006.
[15] FW Dahlquist, RA Elwell, and Peter S Lovely. Studies
of bacterial chemotaxis in defined concentration gradi-
ents. a model for chemotaxis toward l-serine. Journal of
supramolecular structure, 4(3):329–342, 1976.
[16] James A Dix and AS Verkman. Crowding effects on diffu-
sion in solutions and cells. Annu. Rev. Biophys., 37:247–
263, 2008.
[17] David Ellison, Andrew Mugler, Matthew D Brennan,
Sung Hoon Lee, Robert J Huebner, Eliah R Shamir,
Laura A Woo, Joseph Kim, Patrick Amar, Ilya Nemen-
man, et al. Cell–cell communication enhances the capac-
ity of cell ensembles to sense shallow gradients during
morphogenesis. Proceedings of the National Academy of
Sciences, page 201516503, 2016.
[18] Robert G Endres and Ned S Wingreen. Accuracy of
direct gradient sensing by single cells. Proceedings of
the National Academy of Sciences, 105(41):15749–15754,
2008.
[19] Robert G Endres and Ned S Wingreen. Accuracy of di-
rect gradient sensing by cell-surface receptors. Progress
in biophysics and molecular biology, 100(1):33–39, 2009.
[20] Richard A Firtel and Chang Y Chung. The molecular ge-
netics of chemotaxis: sensing and responding to chemoat-
tractant gradients. Bioessays, 22(7):603–615, 2000.
[21] Peter Friedl, Joseph Locker, Erik Sahai, and Jeffrey E
Segall. Classifying collective cancer cell invasion. Nature
Cell Biology, 14(8):777–783, 2012.
[22] Christine Gilles, Myriam Polette, Jean-Marie Zahm,
Jean-Marie Tournier, Laure Volders,
Jean-Michel
Foidart, and Philippe Birembaut. Vimentin contributes
to human mammary epithelial cell migration. Journal of
cell science, 112(24):4615–4625, 1999.
[23] Fran¸cois Graner and James A Glazier. Simulation of
biological cell sorting using a two-dimensional extended
potts model. Physical review letters, 69(13):2013, 1992.
[24] Rama Grantab, Shankar Sivananthan, and Ian F Tan-
nock. The penetration of anticancer drugs through tu-
mor tissue as a function of cellular adhesion and packing
density of tumor cells. Cancer research, 66(2):1033–1039,
2006.
[25] Douglas Hanahan and Robert A Weinberg. The hall-
marks of cancer. cell, 100(1):57–70, 2000.
[26] Douglas Hanahan and Robert A Weinberg. Hallmarks of
cancer: the next generation. cell, 144(5):646–674, 2011.
[27] Bo Hu, Wen Chen, Wouter-Jan Rappel, and Herbert
Levine. Physical limits on cellular sensing of spatial gra-
dients. Physical review letters, 105(4):048104, 2010.
[28] Albertas Janulevicius, Mark van Loosdrecht, and Cris-
tian Picioreanu. Short-range guiding can result in the
formation of circular aggregates in myxobacteria popula-
tions. PLoS Comput Biol, 11, 2015.
[29] Alexandra Jilkine and Leah Edelstein-Keshet. A com-
parison of mathematical models for polarization of single
eukaryotic cells in response to guided cues. PLoS Comput
Biol, 7(4):e1001121–e1001121, 2011.
[30] Alexandre J Kabla. Collective cell migration: leadership,
invasion and segregation. Journal of The Royal Society
Interface, page rsif20120448, 2012.
[31] Kazunari Kaizu, Wiet de Ronde, Joris Paijmans, Koichi
Takahashi, Filipe Tostevin, and Pieter Rein ten Wolde.
The berg-purcell
limit revisited. Biophysical journal,
106(4):976–985, 2014.
[32] Vladimir A Krutovskikh, Colette Piccoli, and Horashi
Yamasaki. Gap junction intercellular communication
propagates cell death in cancerous cells. Oncogene,
21(13):1989–1999, 2002.
[33] DV Krysko, Luc Leybaert, Peter Vandenabeele, and
Katharina D'Herde. Gap junctions and the propaga-
10
tion of cell survival and cell death signals. Apoptosis,
10(3):459–469, 2005.
[34] Il Keun Kwon, Sang Cheon Lee, Bumsoo Han, and Ki-
nam Park. Analysis on the current status of targeted
drug delivery to tumors. Journal of Controlled Release,
164(2):108–114, 2012.
[35] Douglas A Lauffenburger and Alan F Horwitz. Cell mi-
gration: a physically integrated molecular process. Cell,
84(3):359–369, 1996.
[36] Mathias Felix Leber and Thomas Efferth. Molecular prin-
ciples of cancer invasion and metastasis (review). Inter-
national journal of oncology, 34(4):881–895, 2009.
[37] Claire Legrand, Christine Gilles, Jean-Marie Zahm,
Myriam Polette, Anne-C´ecile Buisson, Herv´e Kaplan,
Philippe Birembaut, and Jean-Marie Tournier. Airway
epithelial cell migration dynamics: Mmp-9 role in cell–
extracellular matrix remodeling. The Journal of cell bi-
ology, 146(2):517–529, 1999.
[38] Andre Levchenko and Pablo A Iglesias. Models of eu-
karyotic gradient sensing: application to chemotaxis of
amoebae and neutrophils. Biophysical journal, 82(1):50–
63, 2002.
[39] Pengfei Lu, Andrew J Ewald, Gail R Martin, and Zena
Werb. Genetic mosaic analysis reveals fgf receptor 2 func-
tion in terminal end buds during mammary gland branch-
ing morphogenesis. Developmental biology, 321(1):77–87,
2008.
[40] Oliver J Maclaren, AG Fletcher, HM Byrne, and Philip K
Maini. Models, measurement and inference in epithelial
tissue dynamics. arXiv preprint arXiv:1506.05052, 2015.
[41] Yogeshkumar Malam, Marilena Loizidou, and Alexan-
der M Seifalian. Liposomes and nanoparticles: nanosized
vehicles for drug delivery in cancer. Trends in pharma-
cological sciences, 30(11):592–599, 2009.
[42] Gema Malet-Engra, Weimiao Yu, Amanda Oldani, Javier
Rey-Barroso, Nir S Gov, Giorgio Scita, and Loıc Dupr´e.
Collective cell motility promotes chemotactic prowess
and resistance to chemorepulsion. Current Biology,
25(2):242–250, 2015.
[43] Athanasius F. M. Mar´ee, Veronica A. Grieneisen, and
Paulien Hogeweg. The Cellular Potts Model and Bio-
physical Properties of Cells, Tissues and Morphogenesis.
In Alexander R. A. Anderson, Mark A. J. Chaplain, and
Katarzyna A. Rejniak, editors, Single-Cell-Based Models
in Biology and Medicine, Mathematics and Biosciences in
Interaction. Birkhauser Basel, 2007. DOI: 10.1007/978-
3-7643-8123-3 5.
[44] Roberto Mayor and Carlos Carmona-Fontaine. Keeping
in touch with contact inhibition of locomotion. Trends
in cell biology, 20(6):319–328, 2010.
[45] Colin P McCann, Paul W Kriebel, Carole A Parent, and
Wolfgang Losert. Cell speed, persistence and information
transmission during signal relay and collective migration.
Journal of cell science, 123(10):1724–1731, 2010.
[46] Andrew I Minchinton and Ian F Tannock. Drug penetra-
tion in solid tumours. Nature Reviews Cancer, 6(8):583–
592, 2006.
[47] Andrew Mugler, Andre Levchenko, and Ilya Nemenman.
Limits to the precision of gradient sensing with spatial
communication and temporal integration. Proceedings of
the National Academy of Sciences, page 201509597, 2016.
[48] Carole A Parent. Making all the right moves: chemotaxis
in neutrophils and dictyostelium. Current opinion in cell
biology, 16(1):4–13, 2004.
[49] William J Polacheck, Joseph L Charest, and Roger D
Kamm.
Interstitial flow influences direction of tumor
cell migration through competing mechanisms. Proceed-
ings of the National Academy of Sciences, 108(27):11115–
11120, 2011.
[50] Marten Postma, Jeroen Roelofs, Joachim Goedhart,
Theodorus WJ Gadella, Antonie JWG Visser, and Pe-
ter JM Van Haastert. Uniform camp stimulation of dic-
tyostelium cells induces localized patches of signal trans-
duction and pseudopodia. Molecular biology of the cell,
14(12):5019–5027, 2003.
[51] Alberto Puliafito, Alessandro De Simone, Giorgio Seano,
Paolo Armando Gagliardi, Laura Di Blasio, Federica Chi-
anale, Andrea Gamba, Luca Primo, and Antonio Celani.
Three-dimensional chemotaxis-driven aggregation of tu-
mor cells. Scientific reports, 5, 2015.
[52] William J Rosoff, Jeffrey S Urbach, Mark A Esrick,
Ryan G McAllister, Linda J Richards, and Geoffrey J
Goodhill. A new chemotaxis assay shows the extreme
sensitivity of axons to molecular gradients. Nature neu-
roscience, 7(6):678–682, 2004.
[53] Aida Salameh and Stefan Dhein. Pharmacology of gap
junctions. new pharmacological targets for treatment of
arrhythmia, seizure and cancer? Biochimica et Biophys-
ica Acta (BBA)-Biomembranes, 1719(1):36–58, 2005.
[54] Adrian C Shieh and Melody A Swartz. Regulation of
tumor invasion by interstitial fluid flow. Physical biology,
8(1):015012, 2011.
[55] Jacqueline D Shields, Mark E Fleury, Carolyn Yong, Al-
ice A Tomei, Gwendalyn J Randolph, and Melody A
Swartz. Autologous chemotaxis as a mechanism of tumor
cell homing to lymphatics via interstitial flow and au-
tocrine ccr7 signaling. Cancer cell, 11(6):526–538, 2007.
the advantage
of group navigation. Trends in ecology & evolution,
19(9):453–455, 2004.
[56] Andrew M Simons. Many wrongs:
[57] Loling Song, Sharvari M Nadkarni, Hendrik U Bodeker,
Carsten Beta, Albert Bae, Carl Franck, Wouter-Jan Rap-
pel, William F Loomis, and Eberhard Bodenschatz. Dic-
11
tyostelium discoideum chemotaxis: threshold for directed
motion. European journal of cell biology, 85(9):981–989,
2006.
[58] Maciej H Swat, Gilberto L Thomas, Julio M Belmonte,
Abbas Shirinifard, Dimitrij Hmeljak, and James A
Glazier. Multi-scale modeling of tissues using compu-
cell3d. Methods in cell biology, 110:325, 2012.
[59] A Szab´o, R Unnep, E M´ehes, WO Twal, WS Argraves,
Y Cao, and A Czir´ok. Collective cell motion in endothe-
lial monolayers. Physical biology, 7(4):046007, 2010.
[60] Keila Torres and Susan Band Horwitz. Mechanisms of
taxol-induced cell death are concentration dependent.
Cancer research, 58(16):3620–3626, 1998.
[61] Yuhai Tu. Quantitative modeling of bacterial chemotaxis:
signal amplification and accurate adaptation. Annual re-
view of biophysics, 42:337, 2013.
[62] Pieter Vader, Xandra O Breakefield, and Matthew JA
Wood. Extracellular vesicles: emerging targets for cancer
therapy. Trends in molecular medicine, 20(7):385–393,
2014.
[63] Peter JM Van Haastert and Peter N Devreotes. Chemo-
taxis: signalling the way forward. Nature reviews Molec-
ular cell biology, 5(8):626–634, 2004.
[64] Peter JM Van Haastert and Marten Postma. Biased ran-
dom walk by stochastic fluctuations of chemoattractant-
receptor interactions at the lower limit of detection. Bio-
physical journal, 93(5):1787–1796, 2007.
[65] Shur-Jen Wang, Wajeeh Saadi, Francis Lin, Connie
Minh-Canh Nguyen, and Noo Li Jeon. Differential effects
of egf gradient profiles on mda-mb-231 breast cancer cell
chemotaxis. Experimental cell research, 300(1):180–189,
2004.
[66] Katarina Wolf, Irina Mazo, Harry Leung, Katharina En-
gelke, Ulrich H Von Andrian, Elena I Deryugina, Alex Y
Strongin, Eva-B Brocker, and Peter Friedl. Compensa-
tion mechanism in tumor cell migration mesenchymal–
amoeboid transition after blocking of pericellular prote-
olysis. The Journal of cell biology, 160(2):267–277, 2003.
|
1801.07833 | 2 | 1801 | 2019-05-16T00:49:43 | Effects of controlling parameter on symbolic nonlinear complexity detection | [
"physics.bio-ph",
"physics.data-an"
] | Symbolic transformation, a coarse-graining process, is a crucial prerequisite for and has evidential influence to the symbolic time series analysis. We employ Shannon entropy for a parameter-dependent symbolization, KW (Kurths-Wessel) symbolic method, to test the effects of controlling parameter on its symbolic nonlinear complexity detection. Two chaotic models, logistic and Henon series, and heartbeats of CHF (Congestive Heart Failure) patients, healthy young and elderly subjects from PhysioNet are applied to test the KW symbolic entropy. The complexity-loss theory about aging and diseases in heart rates is validated and reasons that may account for some paradoxes in nonlinear analysis are discussed. Tests results suggest that due to different structural or dynamical properties of different nonlinear systems, controlling parameter of the KW symbolization should be adjusted accordingly to have reliable symbolic nonlinear analysis. | physics.bio-ph | physics |
Effects of controlling parameter on symbolic nonlinear complexity detection
Wenpo Yaoa, Min Wub, Jun Wangc,∗
aSchool of Telecommunications and Information Engineering, Nanjing University of Posts and Telecommunications, Nanjing
cSmart Health Big Data Analysis and Location Services Engineering Lab of Jiangsu Province, Nanjing University of Posts
and Telecommunications, Nanjing 210023, Jiangsu, China
210003, Jiangsu, China
bJinling Hospital, Nanjing 210002, Jiangsu, China
Abstract
Symbolic transformation, a coarse-graining process, is a crucial prerequisite for and has evidential influence
to the symbolic time series analysis. We employ Shannon entropy for a parameter-dependent symbolization,
KW (Kurths-Wessel) symbolic method, to test the effects of controlling parameter on its symbolic nonlinear
complexity detection. Two chaotic models, logistic and Henon series, and heartbeats of CHF (Congestive
Heart Failure) patients, healthy young and elderly subjects from PhysioNet are applied to test the KW
symbolic entropy. The complexity-loss theory about aging and diseases in heart rates is validated and
reasons that may account for some paradoxes in nonlinear analysis are discussed. Tests results suggest that
due to different structural or dynamical properties of different nonlinear systems, controlling parameter of
the KW symbolization should be adjusted accordingly to have reliable symbolic nonlinear analysis.
Keywords:
symbolization, entropy, nonlinear complexity, chaotic model, heartbeat
1. Introduction
Symbolic dynamic deals with robust properties of dynamics without digging into numbers and provides
a rigorous way of looking at real dynamics with finite precision [1]. And the symbolic time series analysis
[2], involving transforming raw time series into symbolic sequence consisting of several discretized symbols,
simplifies the complex system analysis with features of fast calculation, robustness, insensitivity to noise and
so on. At the same time, due to the generally severe degree of the symbolic coarse-graining processes, some
detailed information are lost and the symbolic transformations may have effects on time series analysis [3].
Due to the structural or dynamical differences in different systems, some symbolic transformations have
adjustable controlling parameters to be adaptive and flexible. A symbolic transformation with adjustable
parameter belonging to static methods that transform series into symbolic sequences by dividing series into
finite partitions in the works of Kurths J. and Wessel N. et al. [4, 5, 6] has effective applications in complex
time series analysis. It calculates the time series mean and sets a controlling parameter to transform series
into symbol sequence on the basis of the alphabet A = {0, 1, 2, 3}. In some follow-up researches, scholars find
∗Corresponding author
Email address: [email protected] (Jun Wang)
Preprint submitted to Journal of LATEX Templates
May 17, 2019
it is effective in symbolic nonlinear features detection of physiological signals of cardiac or brain activities
[7, 8, 9].
In these reports, however, there is no in-depth research on the effects of controlling parameter
on symbolic dynamics analysis. The parameter determines the three partitioning lines for the formulation
of symbolic sequence and will affect the nonlinear dynamics extraction. Systematical researches prove that
coarse-graining symbolic methods have effects on the scaling properties of correlated and anti-correlated
signals [3], then how the controlling parameter in KW (Kurths-Wessel) symbolic method affects nonlinear
dynamics complexity detection is what we focus on in this paper.
In our contribution, we employed two chaotic models, logistic and Henon series, and three groups of
heartbeats from PhysioNet to test the effects of the KW symbolization, particularly its controlling parameter,
on nonlinear complexity detection based on Shannon entropy.
2. Methods
2.1. Symbolic Shannon entropy
Symbolic transformation that maps time series into sequence from a given alphabet is the first step
of symbolic time series analysis. The symbolic coarse-graining procedure that some detailed information
is lost while detecting the coarse dynamic behavior [2, 4, 5], therefore, it makes a compromise between
extracting some dynamical information and maintaining sufficient statistics. A pragmatic symbolization is
proposed for physiological time series in contributions of Kurths and Wessel et al.
[4, 5, 6], and it is a
context-dependent transformation having feature of close connection to physiological phenomena and easy
interpretation. Assuming time series X = {x1, x2, . . . , xL}, the KW symbolization transforming X into
symbolic sequence S is carried on as Eq. (1):
si(xi) =
0 : (1 + α)µ < xi < ∞
1 : µ < xi ≤ (1 + α)µ
2 : (1 − α)µ < xi ≤ µ
3 : 0 < xi ≤ (1 − α)µ
(1)
where µ is the series mean and α is special controlling parameter with reference from 0.03 to 0.07, and in the
original literatures analyzing heart rate nonlinear dynamics, α is set to 0.05 [5, 6].
or the convenience of representation, there are negative values in some physiological signals. In some
symbolic works referencing to this method, therefore, time series are divided into positive and negative parts
which are symbolized separately, and the symbolization evolves into Eq. (2):
si(xi) =
0 : (1 + α)µ1 < xi < ∞ or − ∞ < xi < (1 + α)µ2
1 : µ1 < xi ≤ (1 + α)µ1
or
(1 + α)µ2 ≤ xi < µ2
2 : (1 − α)µ1 < xi ≤ µ1
or µ2 ≤ xi < (1 − α)µ2
3 : (1 − α)µ2 ≤ xi ≤ (1 − α)µ1
(2)
where µ1 and µ2 are the means of positive and negative parts.
2
As a comparison, BS (base-scale) symbolization in Eq. (3) of Li and Ning [10, 11], shares some similarities
to the KW symbolic methods that BS method is also a 4-symbol method with α, where µ is the mean of time
series and BS = qPL−1
i=1 [x(i + 1) − x(i)]2/L − 1 represent the root mean square of every two continuous data
difference. To compare with the KW method, we do not consider m-dimensional phase space construction in
the origination of base-scale entropy.
si(xi) =
0 : xi > µ + α × BS
1 : µ < xi ≤ µ + α × BS
2 : µ − α × BS < xi ≤ µ
3 : xi ≤ µ − α × BS
(3)
After symbolization, the next step is to construct m-bit symbolic sequences, leading to a maximum
of 4m different words.
In our following complexity detection of classical chaotic models and real-world
physiological signals, 3-bit encoding is applied to symbolic sequences. There are several statistical approaches
that characterize symbol sequences such as direct visual histograms or quantitative measures based on classical
statistics and information theory [2]. Entropy reflects the static properties and could be applied to original
time series, and it is a useful measure for describing the probabilistic distribution [12]. In our symbolic entropy,
we count probability of each code P (π) = {p(π1), p(π2), . . . , p(πN )} and use classical Shannon entropy for
nonlinear dynamic complexity extraction as Eq. (4), and its normalized form in KW and BS symbolic entropy
is h(m) = H(m)/log4m.
H(m) = −X
i
p(πi)logp(πi), where
p(πi) 6= 0
(4)
3. Chaotic models tests
3.1. Logistic map
The canonical form of logistic difference equation, xn+1 = r · xn(1 − xn), is attractive by the virtue of its
extreme simplicity [13] and is widely applied in chaotic and nonlinear dynamical analysis. The two-degree
polynomial mapping is often referenced as an archetypal example of nonlinear dynamical equations producing
chaotic behaviors [14]. The logistic series (data length is 2000 and x1=0.001) becomes chaotic at the cutting
point r* shown in Fig. 1a and its chaotic behaviors enhance with the increase of r.
KW symbolic entropy with the referenced controlling parameter of 0.05 does not show satisfied results
between r = 3.570 and 3.611 when logistic sequence becomes chaotic shown in Fig. 1b. While when the
controlling parameter is 0.30, the KW symbolic entropy has desirable complexity detections in the whole
regions that it has accurate chaotic complexity captures as shown in Fig. 1c.
Chaotic detection tests show that KW symbolic entropy with α selected from the original interval cannot
provide effectively complexity detections while the symbolic entropy with α chosen close to 0.30 has satisfied
complexity detections in the whole range of logistic map.
3
α=0.05
b
0.6
n
E
W
K
0.5
0.4
0.3
0.2
3.5
r*
3.6
3.7
3.8
3.9
4
r
α=0.20
d
n
E
S
B
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
α=0.30
c
n
E
W
K
0.7
0.6
0.5
0.4
0.3
0.2
3.5
r*
3.6
3.7
3.8
3.9
4
r
0
3.5
r*
3.6
3.7
r
3.8
3.9
4
Figure 1: Logistic equations for varying control parameter from 3.5 to 4, where r*=3.567, exactly 3.569946, is the critical point
for logistic series to step into chaotic sate. a) Bifurcation diagram of logistic map. b) Lyapunov exponent. c) KW symbolic
entropy of α = 0.05. d) BS symbolic entropy of α = 0.20
Similar to KW symbolic entropy, we need to adjust the controlling parameter of base-scale entropy for
chaotic detection of logistic series, and we find the controlling parameter should be chosen around 0.20 shown
in Fig. 1d.
3.2. Henon equations
The Henon map is a classical two-dimensional dissipative quadratic map given by the coupled equations
xn+1 = 1 − α · x2
n + yn, yn+1 = β · xn, and the dynamical system is chaotic for the classical values α = 1.4 and
β = 0.3 [15]. We fix β = 0.3 and set α to 1.33, 1.38 and 1.40 to observe the effects of controlling parameters
on symbolic entropy in the chaotic system. Lyapunov exponents of the three Henon series (data length is
2000 and x1=y1=0.001) are 0.3250, 0.3702 and 0.4077.
a
0.8
n
E
W
K
0.75
0.7
0.65
0.6
0.55
α=0.46
α=0.15
α=1.33
α=1.38
α=1.40
0.75
b
n
E
S
B
0.7
0.65
0.6
0.55
0.5
α=1.33
α=1.38
α=1.40
α=0.50
α=0.70
0.5
0
0.2
0.4
0.6
Controlling parameter α
0.8
1
0.45
0
0.2
0.4
0.6
Controlling parameter α
0.8
1
Figure 2: Symbolic entropy of three Henon outputs with fixed β = 0.3 and α = 1.33, 1.38 and 1.40. a) KW symbolic entropy.
b) Base-scale entropy.
From Fig. 2, the relationships of the three Henon signals' symbolic entropy undergo a variety of changes
as controlling parameter increases from 0.01 to 0.99 with step size of 0.01. Referring to Lyapunov exponents,
the KW symbolic method with α in parameter interval [0.15, 0.45] has rational distinctions. As for base-scale
4
entropy, their controlling parameter also need to be chosen, and intervals [0.51, 0.69] and [0.81, 0.94] have
same chaotic extractions to the KW symbolic method and Lyapunov exponents.
From the nonlinear complexity detections of KW and BS symbolic entropy in the two chaotic models, we
learn that due to the structural and dynamical differences between logistic and Henon systems, controlling
parameter of the symbolic transformation should be adjusted differently to have reliable outcomes.
4. Real-world physiological data tests
4.1. Effects of controlling parameter on symbolic dynamics extraction of heartbeats
Three groups of heartbeats of different subjects, CHF (Congestive Heart Failure) patients [16], healthy
elderly and young people [17] from PhysioBank [18], are applied to test the symbolic entropy. The 15 patients
(11 men aged 22 to 71 and 4 women aged 54 to 63) with severe congestive heart failure underwent about
20 hours data collection. The two healthy groups, 20 young (21 to 34 years old) and 20 elderly (68 to 85
years old) subjects, both have 2 hours of continuous supine resting while continuous data collection. Each
subgroup of the healthy subjects includes equal numbers of men and women. Heart rates derived from ECG
(Electrocardiography) are applied in this section.
Heart rate is typical nonlinear signal with high degree of nonlinear complexity, nonlinearity and nonsta-
tionarity [5, 6, 10, 19, 20]. With controlling parameter increasing from 0.01 to 0.99, the relationships of the
three kinds of heartbeats' KW and BS symbolic entropy are depicted in Fig. 3.
a
n
E
W
K
0.8
0.7
0.6
0.5
0.4
0.3
0
0.2
b
CHF
Elderly
Young
n
E
S
B
0.8
0.7
0.6
0.5
0.4
0.8
1
0.3
0
0.2
CHF
Elderly
Young
0.8
1
0.4
0.6
Controlling parameter α
0.4
0.6
Controlling parameter α
Figure 3: Two symbolic entropy methods of three kinds of heartbeats with increasing controlling parameter. a) KW symbolic
entropy. b) Base-scale entropy.
The choice of parameter is important to three kinds of heartbeats KW nonlinear dynamic analysis in-
dicated by Fig. 3a. Healthy young volunteers heart rates maintain higher KW symbolic entropy than the
elderly ones, and CHF patients heart signals generally have lowest nonlinear complexities. The KW symbolic
entropies of three groups of subjects are divided into three parts:
less than 0.05 is the first interval, from
0.06 to 0.50 is the second part, and more than 0.50 is the third one. KW entropy values of the three groups
heartbeats experience rapid increases to their maximum in the first stage and undergo different degrees of
5
decline and come to their convergences in the second section. Healthy young heart signals complexity con-
verges to around 0.4 when is bigger than 0.40, and the elderly subjects symbolic entropy drop to about 0.36
when controlling parameter increases from 0.15 to 0.20 and bigger. When is bigger than 0.50, KW entropies
of the three groups of heartbeats are convergent and their relationships, the healthy young > the healthy
elderly > the CHF, are stable.
KW symbolic entropy of three kinds of heartbeats with 6 typical parameters from different stages are
chosen for statistically t tests, where 0.03, 0.04 and 0.05 are from the referenced interval in the original
literatures [5, 6], and 0.40, 0.50 and 0.60 are chosen from the convergent stage in Fig. 3a.
Independent
samples t tests for each two heart rates entropy values are listed in Table 1.
Statistical t tests for each two heart rates' KW entropy values are displayed in Table 1. P values of the
CHF and healthy young are all smaller than 0.001 and are not listed in the Table.
Table 1: Three typical relationships of three kinds of HRV ('0.000' should be read as p < 0.001). 0.03, 0.04 and 0.05 are in the
referenced range of the original literatures, and 0.40, 0.50 and 0.60 are chosen from convergent stage
α
0.03
0.04
0.05
0.40
0.50
0.60
CHF-Eld
0.000
0.000
0.006
0.036
0.016
0.011
Eld-Yng
0.023
0.000
0.000
0.013
0.006
0.001
KW symbolic entropy with the above six controlling parameters all have significant differences among
the three kinds of heartbeats statistically (p < 0.05), and results of parameter 0.04 and 0.05 have better
distinctions. With controlling parameter in the referenced interval, however, KW symbolic entropy values
of three kinds of heart signals undergo dramatic changes suggested by Fig. 3a, so complexity detections are
sensitive to the choice of controlling parameter. When is bigger than 0.50, nonlinearity characterized by
symbolic entropy tend to convergent, and parameters in the third stage should be better choices.
The changing trends of base-scale entropy in the three kinds of heartbeats are different from those of
KW symbolic entropy, while the entropy relationships of the different kinds of heart rates are consistent
with the complexity losing theory suggested by Fig. 3b. According to statistical t tests to the distinctions
of the heartbeats BS symbolic entropy, when α < 0.18, the distinctions between of CHF and elderly are
not acceptable (p > 0.05), and when only α >0.47, the separations among the three kinds of heartbeats are
significant statistically.
The complexity-loss theories of aging and disease about heart rates of related literatures [17, 19, 20, 21,
22, 23, 24, 25] are validated in our nonlinear complexity analysis. ECG-derived heartbeat contains valuable
cardiac regulation information and has typical nonlinear dynamical behaviors which attribute to nonlinear
complexity analysis for underlying mechanism of heart activities. Heartbeats of the healthy young generally
yield more dynamical complexity than the elderly ones whose declined integrative cardiac control system
leads to dynamical complexity reduction. Generally, it is accepted that fluctuations patterns of heartbeats
in CHF patients become quite regular and nonlinear dynamics of their heart rates are severely damaged.
6
4.2. Effects of data length on KW symbolic entropy
In characterizing nonlinearity of the three groups of heartbeats, some other symbolic entropy analysis,
such as the base-scale entropy [10, 11], symbolic joint entropy [19] and so forth, have fast calculation and
reliable results even with very short data sets.
In the two chaotic models, we find data length does not
significantly affect results, so in this subsection we conduct research on the effects of data length on the
heartbeats nonlinear complexity detection.
KW symbolic entropy of the three different groups of subjects with data length increasing from 100 to
2900 with step size of 200 are shown in in Fig. 4. We choose controlling parameters of 0.04 and 0.05 in the
referenced interval in the original KW literatures [5, 6] and 0.50 and 0.60 in the convergent stage in Fig. 3a.
W. Yao, M. Wu and J. Wang / Physica A xxx (xxxx) xxx
5. KW symbolic entropy of three groups of heartbeats with the increase of data length.
Figure 4: KW symbolic entropy of three groups of heartbeats with the increase of data length.
of data length on KW symbolic entropy
As can be seen from Fig. 4a and Fig. 4b, when controlling parameters are 0.04 and 0.05, KW symbolic
entropy of three groups of heart signals do not change as data length increases although distinctions between
In characterizeing nonlinearity of the three groups of heartbeats, some other symbolic entropy analysis, such as the
10 11], symbolic joint entropy [19] and so forth, have fast calculation and reliable results even with
In the two chaotic models, we find data length does not significantly affect results, so in this
entropy of CHF heartbeats and that of healthy elderly heart rates deteriorate when data length is 500, and
of 200 are shown in
we conduct research on the effects of data length on the heartbeats' nonlinear complexity detection.
when the length of data sets is longer than 700, statistical differences among the three kinds of heartbeats
KW symbolic entropy of the three different groups of subjects with data length increasing from 100 to 2900 with
. We choose controlling parameters of 0.04 and 0.05 in the referenced interval in the
are significant. In Fig. 4c and Fig. 4d, when data length is less than 1300, complexities of CHF patients
KW literatures [
heartbeats are bigger than those of the healthy elderly which is contradictory to conventional researches, and
KW symbolic entropy of three
of heart signals do not change as data length increases although distinctions between entropy of CHF heartbeats
when data length is 1700 or bigger, the three kinds of heart rates have rational complexity extraction and
of healthy elderly heart rates deteriorate when data length is 500, and when the length of data sets is longer
] and 0.50 and 0.60 in the convergent stage in
As can be seen from
can be distinguished effectively. KW symbolic entropy of the healthy young subjects heart signals maintain
is less than 1300, complexities of CHF patients' heartbeats are bigger than those of the healthy elderly which is
the highest in the four subplots in Fig. 4 and are not affected by data length.
to conventional researches, and when data length is 1700 or bigger, the three kinds of heart rates have
be distinguished effectively. KW symbolic entropy of the healthy young subjects'
From complexity extractions of the three groups of heart rates, we learn that it is more appropriate to
of heartbeats are significant. In
in the four subplots in
by data length.
of the three groups of heart rates, we learn that it is more appropriate to choose
of nonlinear analysis and insensitivity to parameter selection while controlling parameter
to 0.05 is preferred for symbolic transformation if data length is not that large.
7
on symbolic time series analysis and controlling
be adjusted accordingly are shared by related literatures. Xu et al. [ ] demonstrated the effects of
on the scaling behavior of time series analysis. The present authors applied symbolic joint entropy to
in KW and BS symbolic entropy should be adjusted
10
11
12
13
14
15
16
17
18
19
20
21
22
23
choose bigger than 0.50 for the stability of nonlinear analysis and insensitivity to parameter selection while
controlling parameter between 0.03 to 0.05 is preferred for symbolic transformation if data length is not that
large.
Our findings that the symbolic coarse-graining process has effects on symbolic time series analysis and
controlling parameter should be adjusted accordingly are shared by related literatures. Xu et al.
[12]
demonstrated the effects of coarse-graining on the scaling behavior of time series analysis. The present authors
applied symbolic joint entropy to meditation heartbeats and also found that the controlling parameters in
KW and BS symbolic entropy should be adjusted to have reasonable and reliable nonlinear complexity
detection [26]. Also, in a symbolic approach decomposing time series into magnitude and sign series and
reflecting underlying interactions in complex systems by measuring the correlations in the two sub-series by
DFA [27, 28, 29, 30], the scale of DFA or scaling exponents shows effects on the distinction between the
healthy and heart failure heartbeats.
5. Discussions
The symbolic transformation in works of Kurths J. and Wessel N. et al. is analyzed in this paper while
there are still 30 some other issues that need further discussions.
ECG is one of the most direct recordings of cardiac activities. Heartbeat derived from ECG is characterized
with typical nonlinearity while the ECG has clear regularity due to periodic cardiac behaviors. The KW
symbolic entropy for the three kinds of ECG yields some mixed results with the increase of α illustrated by
Fig. 5.
W. Yao, M. Wu and J. Wang / Physica A xxx (xxxx) xxx
6. KW symbolic entropy of CHF, healthy elderly and young ECG.
Figure 5: KW symbolic entropy of CHF, healthy elderly and young ECG.
According to the Fig. 5, the healthy young ECG have lower entropy values than the healthy elderly,
is one of the most direct recordings of cardiac activities. Heartbeat derived from ECG is characterized with typical
to periodic cardiac behaviors. The KW symbolic entropy for the three
and CHF patients ECG symbolic entropy has significantly different changes to the two groups of healthy
of ECG yields some mixed results with the increase of
to the
heartbeats groups which leads to the changes in the relationships of three kinds of ECG KW entropy. And
, the healthy young ECG have lower entropy values than the healthy elderly, and CHF patients'
to the two groups of healthy heartbeats groups which leads to the
in the relationships of three kinds of ECG' KW entropy. And the results are rather confusing that symbolic entropy
of three kinds of heart signals are inconsistent with the aforementioned complexity-loss results in heartbeats. We
the results are rather confusing that symbolic entropy values of three kinds of heart signals are inconsistent
by
with the aforementioned complexity-loss results in heartbeats. We suppose that the contradictory results are
to the different nonlinear features of the heartbeat and ECG.
In other situations with perplexed nonlinear findings, we find there are some following reasons that may be responsible
mainly due to the different nonlinear features of the heartbeat and ECG.
1. Multiscale factor. Multiscale analysis is to construct series
as
/τ
, 1
is scale factor. Some traditional methods are based on single scale, however,
generate nonlinearity over multiple time scale associated with a hierarchy of interacting regulatory
to the fact that the algorithms fail to account for multiple time scales inherent
in physiologic nonlinear dynamics [23 25]. 2. Multi-dimensional phase space. The phase space reconstruction goes as
is embedding dimension and is delay time [31 33]. Low-dimensional
in fact may contain high-dimensional information about complex systems, therefore some information may
be hidden in higher dimensions, and a solution to this situation is to construct multiple dimensional vectors to expose
, where
, . . . ,
= {
8
1)
/τ
1)
10
11
12
13
14
15
16
In other situations with perplexed nonlinear findings, we find there are some following reasons that may
be responsible 9 for possible inconsistent results. 1. Multiscale factor. Multiscale analysis is to construct
coarse-grained series {yj} for the original series {xi}, yτ
i=(j−1)τ +1xi, 1 ≤ j ≤ N/τ where τ is
scale factor. Our symbolic entropy analysis is based on single scale, however, physiologic time series like
j = 1/τ Σjτ
ECG may generate complexity over multiple time scale associated with a hierarchy of interacting regulatory
mechanisms. The paradox maybe due to the fact that the symbolic entropy fails account for multiple time
scales inherent in healthy physiologic dynamics [23, 24, 25]. 2. Multi-dimensional phase space. The space
reconstruction goes as Ym(i) = {x(i), x(i + τ ), . . . , x(i + (m − 1)τ )}, where m is embedding dimension and
τ is delay time [31, 32, 33]. Low-dimensional time series, like ECG, in fact contain all information of the
system, therefore some information may be hidden in higher dimensions and an approach for this situation
is to construct multiple dimensional vectors to expose the hidden structural behaviors. 3. Another reason
contributing to inconsistent results may be that some nonlinear features are inherent in certain forms of the
systems signal and should be preproceed accordingly. The relevant nonlinear information may be reflected
only by some special features extractions or preprocessing approaches. For example, the cardiac nonlinear
information about physiological conditions of aging or diseases are mainly reflected by heartbeats rather than
the ECG. And there might be some other reasons that are responsible for the paradox findings in nonlinear
dynamics analysis remaining unknown.
In our nonlinear analysis, Shannon entropy is employed for the symbolic sequences. Entropy measures
have practical and theoretical significance in applications of complexity detections of chaotic or real-world
time series. In a systematic study on the entropy methods [12], entropy measures are proved to be able to
characterize the heartbeats in different physiological and clinical states and entropy is a suitable parameter
in measuring the variability and preferable to be applied to the original time series. Symbolization is the
first step in symbolic time series analysis, and it can be followed by a variety of different analytical methods,
such as classical statistics, entropy, time irreversibility and other measures [2, 20, 34]. Different parameters
target on different aspects of information, therefore in applying other analytical measures, it need to find
which controlling parameter has the optimal outcome by more comprehensive researches.
6. Conclusions
Through the KW symbolic entropy analysis in chaotic models and real-world heartbeats, we learn that the
symbolic method should adjust its controlling parameter to achieve reliable nonlinear complexity detections
because of different structural or dynamical properties of nonlinear systems. And we further verify the
complexity-loss theory about aging and diseased cardiac regulation system.
7. Acknowledgments
The work is supported by the National Natural Science Foundation of China (Grant Nos. 31671006,
61771251), Jiangsu Provincial Key R & D Program (Social Development) (Grant No.BE2015700, BE2016773),
9
Natural Science Research Major Programmer in Universities of Jiangsu Province (Grant No.16KJA310002).
References
References
[1] B. L. Hao, Symbolic dynamics and characterization of complexity, Physica D Nonlinear Phenomena
51 (1-3) (1991) 161 -- 176.
[2] C. S. Daw, C. E. A. Finney, E. R. Tracy, A review of symbolic analysis of experimental data, Review of
Scientific Instruments 74 (2) (2003) 915 -- 930.
[3] Y. Xu, Q. D. Ma, D. T. Schmitt, P. Bernaolagalvn, P. C. Ivanov, Effects of coarse-graining on the
scaling behavior of long-range correlated and anti-correlated signals, Physica A statistical Mechanics &
Its Applications 390 (23-24) (2011) 4057.
[4] J. Kurths, A. Voss, P. Saparin, A. Witt, H. J. Kleiner, N. Wessel, Quantitative analysis of heart rate
variability, Chaos: An Interdisciplinary Journal of Nonlinear Science 5 (1) (1995) 88 -- 94.
[5] N. Wessel, C. Ziehmann, J. Kurths, U. Meyerfeldt, A. Schirdewan, A. Voss, Short-term forecasting of
life-threatening cardiac arrhythmias based on symbolic dynamics and finite-time growth rates, Physical
Review E 61 (1) (2000) 733 -- 739.
[6] N. Wessel, H. Malberg, R. Bauernschmitt, J. Kurths, Nonlinear methods of cardiovascular physics and
their clinical applicability, International Journal of Bifurcation & Chaos 17 (10) (2007) 3325 -- 3371.
[7] Y. Wenpo, W. Jun, Multi-scale symbolic transfer entropy analysis of eeg, Physica A 484 (2017) 276 -- 281.
[8] Y. Wang, F. Z. Hou, J. F. Dai, X. F. Liu, J. Li, J. Wang, Analysis on relative transfer of entropy based
on improved epileptic eeg, Acta Physica Sinica 63 (21) (2014) 218701.
[9] Y. Wang, F. Z. Hou, J. F. Dai, X. F. Liu, J. Li, J. Wang, Time irreversibility analysis of ecg based on
symbolic relative entropy, Acta Physica Sinica 60 (11) (2011) 25092515.
[10] J. Li, X. Ning, Dynamical complexity detection in short-term physiological series using base-scale en-
tropy., Physical Review E 73 (1) (2006) 88 -- 101.
[11] J. Li, X. Ning, Classification of physiologic and synthetic heart rate variability series using base-scale
entropy, Physica A Statistical Mechanics & Its Applications 384 (2) (2007) 423 -- 428.
[12] W. Xiong, L. Faes, P. C. Ivanov, Entropy measures, entropy estimators, and their performance in
quantifying complex dynamics: Effects of artifacts, nonstationarity, and long-range correlations, Physical
Review E 95 (6) (2017) 062114.
10
[13] R. M. May, Simple mathematical models with very complicated dynamics, Nature 261 (5560) (1976)
459 -- 467.
[14] G. Boeing, Visual analysis of nonlinear dynamical systems: Chaos, fractals, self-similarity and the limits
of prediction, Systems 4 (4) (2016) 37.
[15] M. Henon, A two-dimensional mapping with a strange attractor, Communications in Mathematical
Physics 50 (1) (1976) 69 -- 77.
[16] D. S. Baim, W. S. Colucci, E. S. Monrad, H. S. Smith, R. F. Wright, A. Lanoue, D. F. Gauthier, B. J.
Ransil, W. Grossman, E. Braunwald, Survival of patients with severe congestive heart failure treated
with oral milrinone, Journal of the American College of Cardiology 7 (3) (1986) 661 -- 670.
[17] N. Iyengar, C. K. Peng, R. Morin, A. L. Goldberger, L. A. Lipsitz, Age-related alterations in the fractal
scaling of cardiac interbeat interval dynamics., American Journal of Physiology 271 (4 Pt 2) (1996)
R1078.
[18] A. L. Goldberger, L. A. N. Amaral, L. Glass, J. M. Hausdorff, P. C. Ivanov, R. G. Mark, J. E. Mietus,
G. B. Moody, C. K. Peng, H. E. Stanley, Physiobank, physiotoolkit, and physionet, Circulation 101 (23)
(2000) 215 -- 220.
[19] Y. Wenpo, W. Jun, Double symbolic joint entropy in nonlinear dynamic complexity analysis, Aip Ad-
vances 7 (7) (2017) 075313.
[20] W. Yao, W. Yao, J. Wang, J. Dai, Quantifying time irreversibility using probabilistic differences between
symmetric permutations, Physics Letters A 383 (8) (2019) 738 -- 743.
[21] P. C. Ivanov, L. A. N. Amaral, A. L. Goldberger, S. Havlin, M. G. Rosenblum, Z. R. Struzik, H. E.
Stanley, Multifractality in human heartbeat dynamics, Nature 399 (6735) (1999) 461 -- 465.
[22] A. L. Goldberger, C. K. Peng, L. A. Lipsitz, What is physiologic complexity and how does it change
with aging and disease?, Neurobiology of Aging 23 (1) (2002) 23 -- 26.
[23] M. D. Costa, A. L. Goldberger, C. K. Peng, Multiscale entropy analysis of complex physiologic time
series., Physical Review Letters 89 (6) (2002) 068102.
[24] M. D. Costa, A. L. Goldberger, C. K. Peng, Multiscale entropy analysis of biological signals., Physical
Review E 71 (2 Pt 1) (2005) 021906.
[25] M. D. Costa, C. K. Peng, A. L. Goldberger, Multiscale analysis of heart rate dynamics: Entropy and
time irreversibility measures, Cardiovascular Engineering 8 (2) (2008) 88 -- 93.
[26] Y. Wenpo, Z. Yuping, W. Jun, Quantitative analysis in nonlinear complexity detection of meditative
heartbeats, Physica A 7 (7) (2018) 075313.
11
[27] Y. Ashkenazy, P. C. Ivanov, S. Havlin, C. K. Peng, A. L. Goldberger, H. E. Stanley, Magnitude and
sign correlations in heartbeat fluctuations, Physical Review Letters 86 (9) (2001) 1900 -- 1903.
[28] Y. Ashkenazy, S. Havlin, P. C. Ivanov, C. K. Peng, V. Schulte-Frohlinde, H. E. Stanley, Magnitude and
sign scaling in power-law correlated time series, Physica A: Statistical Mechanics and its Applications
323 (5) (2003) 19 -- 41.
[29] P. C. Ivanov, Q. D. Ma, R. P. Bartsch, J. M. Hausdorff, L. A. N. Amaral, V. Schulte-Frohlinde, H. E.
Stanley, M. Yoneyama, Levels of complexity in scale-invariant neural signals, Physical Review E 79 (4)
(2009) 041920.
[30] M. Gomez Extremera, P. Carpena, P. C. Ivanov, P. A. Bernaola-Galv´an, Magnitude and sign of long-
range correlated time series: Decomposition and surrogate signal generation, Physical Review E 93 (4)
(2016) 042201.
[31] M. Casdagli, S. Eubank, J. D. Farmer, J. Gibson, State space reconstruction in the presence of noise,
Physica D Nonlinear Phenomena 51 (1-3) (1991) 52 -- 98.
[32] H. S. Kim, R. Eykholt, J. D. Salas, Nonlinear dynamics, delay times, and embedding windows, Physica
D Nonlinear Phenomena 127 (12) (1999) 48 -- 60.
[33] L. Cao, Practical method for determining the minimum embedding dimension of a scalar time series,
Physica D nonlinear Phenomena 110 (1-2) (1997) 43 -- 50.
[34] W. Yao, W. Yao, J. Wang, J. Dai, Quantifying time irreversibility using probabilistic differences between
symmetric permutations, Physics Letters A 383 (8) (2019) 738 -- 743.
12
|
1805.03197 | 1 | 1805 | 2018-05-07T19:59:50 | A guide to emerging technologies for large-scale and whole brain optical imaging of neuronal activity | [
"physics.bio-ph",
"physics.optics"
] | The mammalian brain is a densely interconnected network that consists of millions to billions of neurons. Decoding how information is represented and processed by this neural circuitry requires the ability to capture and manipulate the dynamics of large populations at high speed and resolution over a large area of the brain. While there has been a rapid increase in use of optical approaches in the neuroscience community over the last two decades, most microscopy approaches lack the ability to record the activity of all neurons comprising a functional network across the mammalian brain at relevant temporal and spatial resolution. In this review, we survey the recent development in the optical calcium imaging technologies in this regard and provide an overview of the strengths and limitations of each modality and their potential for scalability. We provide a guidance from a biological user perspective that is driven by the typical biological applications and sample conditions. We also discuss the potential for future advances and synergies that could be obtained through hybrid approaches or other modalities. | physics.bio-ph | physics | A guide to emerging technologies for large-scale and whole brain optical
imaging of neuronal activity
Siegfried Weisenburgera, Alipasha Vaziria,b,c,∗
aLaboratory of Neurotechnology and Biophysics, The Rockefeller University, New York, NY, USA
bKavli Neural Systems Institute, The Rockefeller University, New York, NY, USA
c Research Institute of Molecular Pathology, Vienna, Austria
8
1
0
2
y
a
M
7
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
9
1
3
0
.
5
0
8
1
:
v
i
X
r
a
Abstract
The mammalian brain is a densely interconnected network that consists of millions to billions of
neurons. Decoding how information is represented and processed by this neural circuitry requires the
ability to capture and manipulate the dynamics of large populations at high speed and resolution
over a large area of the brain. While there has been a rapid increase in use of optical approaches in
the neuroscience community over the last two decades, most microscopy approaches lack the ability
to record the activity of all neurons comprising a functional network across the mammalian brain at
relevant temporal and spatial resolution.
In this review, we survey the recent development in the optical calcium imaging technologies in this
regard and provide an overview of the strengths and limitations of each modality and their potential
for scalability. We provide a guidance from a biological user perspective that is driven by the typical
biological applications and sample conditions. We also discuss the potential for future advances and
synergies that could be obtained through hybrid approaches or other modalities.
Keywords: Ca2+ imaging, Volumetric imaging, large-scale, neural circuit dynamics, high-speed,
functional network
1. Introduction
One of the key goals of neuroscience is to deci-
pher how the complex dynamics of neuronal activ-
ity in the brain is related to perception, cognition
and behavior, and what the underlying principles
for information processing by these circuits are
(Averbeck et al. 2006). Mammalian brains con-
sist of millions to billions of neurons, each of which
making thousands of connections to other neurons
(Herculano-Houzel et al. 2006, 2007). This re-
sults in an extreme density of interconnectivity in
the brain, so that randomly selected neurons are,
on average, only via a few neurons connected to
∗Corresponding author.
Email addresses: [email protected]
(Siegfried Weisenburger), [email protected]
(Alipasha Vaziri)
each other. From this anatomical fact alone it be-
comes clear that, in such a densely interconnected
system, information can be often represented in a
distributed fashion and any given neuron can be
involved in representation of multiple stimuli or
behavioral states in a context-dependent manner.
Progress in identifying the "neuronal code" has
been limited by the lack of appropriate tools and
technologies that would allow to record and ma-
nipulate the activity of many - and potentially
all - neurons within a functional network. This
requires methods that enable: (1) control of ar-
bitrary spatiotemporal patterns of neural activity
and (2) read out of the entire system at adequate
temporal and spatial resolution.
One of the first approaches for simultaneous
measurement of the activity of multiple neurons
May 10, 2018
has been based on electrodes that capture the
electrical responses from neurons within the brain
region of interest (Krüger 2005). Extracellular
neuronal recordings with multi-electrode arrays
(Buzsaki 2004) are now used together with algo-
rithms for spike sorting to distinguish signals from
different neurons based on their waveform (Harris
et al. 2016, Rey et al. 2015). However, insertion
of electrodes is invasive and approaches for spike
sorting work only within the immediate vicinity
of the electrodes where the peak amplitudes are
high. This, depending on the number of elec-
trodes used, currently limits the effective number
of neurons that can be interrogated to ~1,000 –
2,000 neurons (Berenyi et al. 2014, Hilgen et al.
2017).
Functional MRI (Ogawa et al. 1990), in con-
trast, allows for whole-brain recordings in a va-
riety of nervous systems including humans in a
non-invasiveness manner. The contrast in fMRI
stems from the hemodynamic response, which is
based on changes of the blood oxygenation level
in response to neuronal activity. Consequently,
fMRI is not a direct measure of neuronal activity
and provides a temporal resolution that is only
in the sub-Hz regime. Moreover, limited by the
gradient of magnetic fields that can be created,
its spatial resolution is limited to the millimeter
range such that each voxel contains millions of
neurons (Huettel et al. 2008).
Another approach to record neuronal activity
is based on optical microscopy in combination
with functional contrast agents. The first exper-
iments based on this approach used purified pro-
teins from jellyfish (Shimomura et al. 1962), and
organic fluorescent dye molecules that are voltage
sensitive as reporters of membrane depolarization,
as well as dyes that can act as reporters of con-
centration changes of ions such as Ca2+ (Grinvald
et al. 1987, Grynkiewicz et al. 1985, Salzberg et
al. 1977). Transformative for the field of opti-
cal imaging of neuronal activity was the concep-
tion of GECIs (Miyawaki et al. 1997, Nakai et
al. 2001). The advent of these novel probes al-
lowed advanced optical microscopy methods such
as 2p excitation fluorescence microscopy (Denk et
al. 1990, So et al. 2000) to be used to their full po-
tential, allowing to record neuronal activity at cel-
lular to sub-cellular resolution at tissue depths up
to about 1mm, comparable to the thickness of the
mouse cortex (Chen et al. 2013, Mao et al. 2008).
In the past decade, the number of new tools and
technologies has soared, with an exponential in-
crease over the last five years, in part fueled by
interdisciplinary ventures such as the Brain Re-
search through Advancing Innovative Neurotech-
nologies (BRAIN) Initiative (Insel et al. 2013)
as well as other initiatives (Yuste & Bargmann
2017). Today, whole-brain functional imaging of
small organisms such as C. elegans (Schrödel et
al. 2013) or zebrafish larvae (Ahrens et al. 2013,
Prevedel et al. 2014) have been demonstrated,
and current developments are aimed at large-scale
recording and interrogation of mammalian brains
with ever-increasing neuronal population size.
The aim of this review is to provide an overview
of the various microscopy techniques for Ca2+
imaging, that are available today, from a concep-
tual point of view while providing guidelines for
biological users which allow them to identify the
most-suited techniques for their particular biolog-
ical questions and context. While we have lim-
ited ourselves to only optical methods for Ca2+
imaging in this review, a number of the discussed
tradeoffs also apply to optogenetics (Boyden et al.
2005, Nagel et al. 2002, Vaziri & Emiliani 2012,
Zhang et al. 2007) and optical voltage imaging
(Xu et al. 2017).
2. General microscopy concepts and con-
siderations
To gain access to the full level of information
that a brain represents at any given moment in
time, a method capable of capturing the activ-
ity of every single neuron with adequate tempo-
ral and synaptic resolution over the entire brain
would be required.
It would need to allow for
non-invasive recording in mammalian brains over
extended periods of time, typically on the order
of hours as needed for behavioral studies, and
for a number of sessions over several weeks to
months. Figure 1 gives an overview of some nu-
merical values of the key anatomical features for
May 10, 2018
the brains of traditionally used model systems,
and it illustrates the order of magnitude differ-
ences in neuron numbers, synapses or brain vol-
ume, that imaging methods must cover. While
none of the currently available techniques is able
to achieve these requirements at the same time,
a plethora of tools, especially various microscopy
approaches, have emerged over the recent years
aimed at addressing this grand challenge at differ-
ent levels.
In particular, volumetric microscopy
methods have been shown to achieve single cell
resolution - and in some cases synaptic resolu-
tion - at acquisition rates comparable to or faster
than the average rate of activity of cortical neu-
rons.
The anatomical properties (Fig. 1) define the
required spatial and temporal resolution, how-
ever, they also establish the sparsity in time and
space, within which methods need to operate, but
also what they can actively exploit to relax some
of the optical constraints. Key parameters of an
imaging modality are its spatial resolution, the ac-
quisition speed and the volume size it can record
from. These parameters are linked and further de-
termined by biological and technical limitations:
When imaging in a transparent medium, the NA
and the wavelength λ determine the spatial reso-
lution, d = λ/(2NA) (Weisenburger & Sandogh-
dar 2015). In diffraction-limited imaging, the spa-
tial resolution d, the size of the acquisition volume
V, and the number of acquired voxels n per time
unit determine the volume rate. Neglecting any
overhead due to technical limitations, the number
of acquired voxels per second is n = M /∆t, where
∆t is the voxel dwell time and M is the multi-
plicity of the acquisition, that is the number of
voxels recorded in parallel. To avoid acquisition
crosstalk and obtain optimal signal per voxel in
fluorescence imaging, the dwell time ∆t must be
longer than the fluorescence lifetime, which is the
time the fluorophore spends in the excited state
before emitting a fluorescence photon, typically
on the order of a few nanoseconds. Besides this
fundamental limit, properties of the sample and
the used fluorophore, such as fluorophore concen-
tration and quantum yield, limit the dwell time
in a more practical fashion in the required SNR.
A limiting factor for all imaging modalities is
photodamage: One source of damage is related
to the average amount of power that is deposited
into the tissue causing tissue heating. Onset of
damage has been observed at temperatures above
40◦C in the mammalian brain (Podgorski & Ran-
ganathan 2016). At the same time it has been
shown that cranial windows and immersion wa-
ter can lower the superficial brain temperature
to 32◦C (Kalmbach & Waters 2012). The maxi-
mum average laser power for which heating will be
non-detrimental depends on the illuminated vol-
ume, the size of the window and imaging depth,
as well as the thermal properties of the brain tis-
sue and the vasculature. For 2p in vivo imag-
ing in the mouse brain, average laser powers of
about 250 mW have been used without observa-
tion of damage as reported by immunohistochem-
istry (Podgorski & Ranganathan 2016, Prevedel
et al. 2016). Another possible source of dam-
age is nonlinear photo-damage which is caused
by nonlinear tissue absorption, intracellular di-
electric breakdown or possible plasma formation
due to high peak intensities (Hopt & Neher 2001,
Koester et al. 1999). Reported damage thresh-
olds for non-linear photo-damage are ~20 nJ/µm2
(Hopt & Neher 2001), which corresponds to ~10
mW in a diffraction-limited focus of a standard
80-MHz titanium sapphire oscillator. Any limi-
tations in the employable average power or peak
intensity affects the available signal or SNR and
has, in turn, an influence on reachable depth or
acquisition speed.
Since the discussed tissue heating and photo-
damage mechanisms restrict the maximum aver-
age power and peak intensities that the issue can
tolerate, they also indirectly restrict the multi-
plicity M,
the number of voxels that are simul-
taneously recorded at each time point. Further-
more, when imaging in scattering tissue while us-
ing a 2D detector array such as a camera, the
ability to distinguish nearby voxels degrades with
increasing depth due to scatter-induced crosstalk
between neighboring voxels, setting another limit
on the practically utilizable multiplicity M. These
sets of relations and interdependencies between
the fundamental imaging parameters and condi-
May 10, 2018
Figure 1: Overview of numerical values for the key anatomical parameters of the nervous systems /brains of tradi-
tionally used model organisms in neuroscience. The table illustrates the order of magnitude differences in neuron
numbers, synapses, and brain volumes that imaging methods need to cover. Numbers based on (White et al. 1986): C.
elegans,(Naumann et al. 2010): zebrafish larvae, (Gouwens & Wilson 2009, Rein et al. 2002): drosophila,(Herculano-
Houzel et al. 2006, Howarth et al.2012, Kovacevic 2004, Simpson 2009): mouse, (Braitenberg & Schüz 1998, Buzsaki
&Mizuseki 2014, Herculano-Houzel et al. 2006, Martin et al. 2010, Sahin et al.2001): rat, (Buzsaki & Mizuseki 2014,
Herculano-Houzelet al. 2007): marmoset.
tions, within which samples can be imaged, shows
that there is no single ideal method. Each system
or biological question brings a different set of con-
ditions, challenges and opportunities.
3. Molecular indicators for optical readout
of neuronal activity
Fluorescence is a key contrast mechanism in bi-
ological imaging due to the diversity of labeling
strategies that it offers, its high contrast, sensitiv-
ity, and the possibility to obtain cell type specific
labeling when genetically expressible fluorescence
labels are used. In the case of functional imaging,
neuronal activity is either directly or indirectly
translated into the modulation of a fluorescence
signal by a molecular sensor. A dramatic advance
for functional fluorescence imaging was the advent
of GECIs in the beginning of the 2000s (Miyawaki
et al. 1997, Nakai et al. 2001). GECIs revolu-
tionized functional neuronal imaging for several
reasons: The most recent generation of GECIs
provide superb SBR (Chen et al. 2013) which
has enabled recording of neuronal activity at sev-
eral hundreds of microns in the scattering rodent
tissue in vivo. Additionally, they offer genetic
targeting, allowing for recording of activity from
specific neuronal cell types. Finally, they also
allow for a diverse range of delivery strategies.
Viral transduction strategies through e.g.
lenti-
and adeno-associated viruses (Dittgen et al. 2004,
Tian et al. 2009) including systemic delivery ap-
proaches (Chan et al. 2017, Foust et al. 2008,
Shimogori & Ogawa 2008), as well as methods for
genomic integration of the transgenes and gener-
ation of transgenic animals (Ji et al. 2004, Tallini
et al. 2006, Zariwala et al. 2012) are now being
routinely used (Grienberger & Konnerth 2012).
The Ca2+ reporter GCaMP was created by fus-
ing a GFP to calmodulin, a Ca2+ binding protein,
such that a conformational change modifies the
chemical environment of the chromophore upon
Ca2+ binding. Thus, neurons expressing GCaMP
show a weak fluorescence signal when they are in-
active and become brightly fluorescent upon on-
set of neuronal activity due to the influx of Ca2+
ions. Ca2+ reporters have been continuously im-
proved in brightness, response time, fluorescence
modulation or binding affinity (Akerboom et al.
2012, Chen et al. 2013, Nagai et al. 2004, Tian
et al. 2009). The currently widely used version
GCaMP6f is one of the fastest Ca2+ indicators
with a decay time of ~140 ms and is still very
bright with a strong fluorescence modulation of
May 10, 2018
~20 % for a single action potential (Chen et al.
2013).
All Ca2+ imaging techniques share some funda-
mental limitations due to the fact that they are
a second messenger for the membrane depolar-
ization (Lin & Schnitzer 2016). The time scale
of the fluorescent signal in Ca2+ imaging is dic-
tated by diffusion, Ca2+ buffering and cooperativ-
ity. Their reaction is about two orders of magni-
tude slower than the time scale of an action poten-
tial which results in inaccuracies and variabilities
in spike timing and response characteristics. The
Ca2+ signal also saturates during bursts of activ-
ity, limiting the maximal fluorescence modulation
to ~200 %. Furthermore, GECIs typically cannot
report any membrane hyperpolarization or sub-
threshold voltage changes (Lin & Schnitzer 2016).
Thus, parallel efforts are aimed at the de-
velopment of genetically encoded voltage indica-
tors (GEVIs) as a direct reporter of the mem-
brane polarization dynamics. Currently, GEVIs
do not provide the same activity contrast ratios as
GECIs, but they are able to capture neuronal sig-
nals with ~10 ms time resolution (Chamberland
et al. 2017, Hochbaum et al. 2014). We expect
that their further development will lead to a syn-
ergetic impact with the technological progress on
the microscopy side as discussed in this review.
4. Computational signal extraction and
computational imaging
The steep increase in computer processing
power over the past decades enabled the emer-
gence of computational imaging methods where
computation becomes an integral part of the
imaging process.
In a range of computational
imaging strategies, any form of prior knowledge
about the sample or the experimental instrumen-
tation is used to enable or enhance the capabil-
ities of the imaging system. Examples include
statistical tools such as independent component
analysis (ICA) (Mukamel et al. 2009) or non-
negative matrix factorization (NMF) techniques
(Pnevmatikakis et al. 2016) that have been shown
to allow the extraction and de-mixing of neuronal
signals even when they are overlapping in space
or time. While ICA is a linear de-mixing method,
nonlinear, NMF-based techniques allow for a bet-
ter de-mixing in the case of strongly overlapping
sources (Pnevmatikakis et al. 2016). NMF makes
use of the fact that spatiotemporal activity can be
approximated as the product of a spatial matrix
containing the location of each neuron and a tem-
poral matrix representing their activity. When
seeded or informed by prior knowledge about the
location of neurons, such an approach can de-mix
neuronal signals stemming from deep inside scat-
tering tissue even when combined with wide-field
detection, as recently demonstrated in combina-
tion with LFM by the technique of seeded itera-
tive de-mixing (SID) (Nöbauer et al. 2017).
In general, computational deconvolution and
reconstruction methods can reduce the burden on
the optics or generally the demands on instru-
mentation (Friedrich et al. 2017). Extrapolat-
ing from the current trends that point towards an
ever-increasing complexity of the imaging appara-
tus, the use of computational imaging and other
tools to relax the constraints on the instrumen-
tation are expected to become increasingly im-
portant and an integral aspect of future optical
engineering efforts.
5. Categorization of current microscopy
methods used for Ca2+ imaging
Current techniques for Ca2+imaging can be
broadly categorized according to their degree of
parallelization of acquisition and whether the ac-
quisition is unbiased or prior-based as the obtain-
able acquisition speeds and the applicability of
the various methods to different scattering prop-
erties of the sample are determined by them (Fig.
2). In addition, there are currently several strate-
gies and emerging techniques for optical recording
at greater depths, which we will discuss in this
section and describe their advantages and limita-
tions.
5.1. Mode of acquisition: Parallel versus
sequential
In a sequential acquisition, a focal spot is
scanned across one or more dimensions in space to
May 10, 2018
Figure 2: Illustration of imaging modalities, categorized based on the acquisition mode (fully sequential, 1D parallel,
2D parallel and 3Dparallel) and sampling strategy (unbiased /prior-based sampling). Excitation and detection is in
epi-configuration from the top if not indicated otherwise. See legend for an explanation of the used symbols.
cover the entire volume. In a parallel acquisition
mode, a certain number or all voxels are recorded
simultaneously.
Based on this definition, a range of techniques
exists that share some degree of parallelization
with some degree of sequential acquisition (Fig 2).
In fully sequential techniques such as diffraction-
limited point scanning, the location of the sig-
nal is determined by the instantaneous position of
the excitation beam. Consequently, a point detec-
tor can be used that collects emitted fluorescence
photons irrespective of the path on which they
reach the detector. Thus, fully sequential acqui-
sition techniques, and to a lesser extent partially
sequential techniques, provide robustness towards
scattering and are therefore well-suited for deep
tissue imaging. They typically also deposit only
a small amount of energy per time unit in the
sample which reduces or avoids the risk of tissue
damage. Yet, this comes at the cost of a reduced
time resolution. A parallelized acquisition mode
based on a detector array or a camera allows for
higher volume rates, but, depending on the scat-
tering properties of the tissue, at the cost of po-
tential scatter-induced crosstalk between neigh-
boring camera pixels. Furthermore, the deposited
average power and the potential risk for tissue
damage increases with the degree of paralleliza-
May 10, 2018
tion.
5.1.1. Sequential acquisition:
The most basic fully sequential technique for
3D imaging using point scanning is confocal LSM
(Pawley 2006), Fig. 2a. Confocal microscopy is
rarely used for in vivo imaging in scattering brains
due its low efficiency to collect emitted fluores-
cence photons and level of fluorophore bleaching.
As a result, the application of some of its vari-
ants such as confocal spinning disk microscopy
has been limited to small and semi-transparent
organisms such as C. elegans (Kato et al. 2015).
A prominent technique for functional imaging
using point scanning is 2p LSM (Denk et al. 1990,
So et al. 2000). Here, the absorption of two lower-
energy photons is necessary for the fluorophore to
be able to emit fluorescence. This means that the
2p absorption probability depends on the light in-
tensity in the focal spot squared. To reach the
necessary peak intensities for 2p fluorescence ex-
citation, pulsed femtosecond lasers are typically
utilized. The nonlinearity in the excitation comes
with another benefit: Only fluorescent molecules
in the laser focus experience a sufficient inten-
sity for 2p excitation and subsequent fluorescence
emission. This results in a significant reduction
of the out-of-focus fluorescence and thereby an in-
crease in the SNR and SBR while minimizing pho-
tobleaching outside of the focus. 2p excitation is
also beneficial for a deeper penetration of the exci-
tation light owing to the lower absorption of tissue
in the NIR region of the spectrum and because the
scattering intensity approximated by Mie scatter-
ing scales as ~1/λ4. An additional design consid-
eration when using ultrafast laser pulses for ex-
citation in both scattering and transparent tissue
is pulse broadening due to dispersion which re-
duces the fluorophore excitation probability and
therefore the SNR (Trebino & Zeek 2000).
Although 2p microscopy has enabled a wide
range of studies in scattering brain tissue, pri-
marily in rodents, during the past two decades,
its low speed has limited neuroscience studies to
planar recordings, even with fast resonant scan-
ning schemes (Kerr & Denk 2008).
One way to increase the speed is to intro-
duce multiplicity in the acquisition. This can be
done by scanning across the volume with multiple
beams simultaneously, a technique termed tem-
poral multiplexing (Amir et al. 2007, Cheng et
al. 2011, Stirman et al. 2016). Here, the exci-
tation laser beam is divided into multiple beam-
lets, which are delayed with respect to each other,
and then are used for simultaneous excitation and
recording in different lateral or axial positions
in the imaging volume (Fig. 2b). While a sin-
gle point detector such as a PMT is used, it is
possible in this scheme to distinguish the fluores-
cence signal from different spatial regions by their
different arrival times at the PMT. Four-times
multiplexing schemes have been demonstrated, in
which the degree of multiplicity was limited by
the typical 80 MHz laser repetition rate of a tita-
nium sapphire oscillator, corresponding to a 12.5
ns inter-pulse interval, and the fluorescence life-
time which is on the order of a few nanoseconds
(Cheng et al. 2011). Using this approach, multi-
plane diffraction-limited imaging over a FOV of
400 x 400 µm has been shown for four planes at a
rate of 60 fps over an axial range of less than 100
µm (Cheng et al. 2011).
In the above scheme,
the required average power scales linearly with
the number of multiplexed beams, so that pho-
todamage due to tissue heating ultimately limits
the multiplicity that can be practically achieved
using this technique. This method is also not eas-
ily scalable as the complexity of the optical setup
increases with the multiplicity.
Another approach to increase the volumetric
recording speed in sequential imaging is to op-
timize and adapt the spatial resolution and sam-
pling frequency to the structure of interest (Fig.
2c). Since the average size of neuronal cell bodies
in mammalian cortex (~10-20 µm) is more than
ten times larger in diameter than a diffraction-
limited focus, spatial resolution can be traded
for speed, allowing recording of neuronal signals
from larger populations at a higher speed while
maintaining single-cell resolution, and depending
on excitation depth can allow for a higher SNR
at the same average power. Fundamental con-
straints from optics do not allow an arbitrarily
shaped excitation PSF when a Gaussian beam is
May 10, 2018
Figure 3: Overview of the categories and key imaging properties of the available techniques for functional optical imaging,
and their strengths and limitations. The modalities are categorized based on the acquisition mode(fully sequential,
1D parallel, 2D parallel and 3D parallel) and sampling strategy (unbiased/prior-based sampling). The key imaging
properties, acquisition speed, scattering compatibility, suitability for awake imaging(including susceptibility to motion
artifacts), photodamage, optical access, and scalability of the method, are rated with ++ (best, green), + (green), 0
(yellow), − (red), −− (worst, red). Example references for the techniques are provided.
Figure 4: Overview of excitation modalities and their cor-
responding PSF of different imaging method: (a) Gaussian
focus, (b) under-filled Gaussian focus, (c) Isotropic focus
with light sculpting using temporal focusing, (d) Bessel
beam generated by an annular pattern in the back aper-
ture, (e) V-shaped PSF.
used.
In particular, isotropic resolution cannot
be achieved with a Gaussian beam because of the
intrinsic coupling of the lateral (d) and the axial
(z) confinement of excitation in a Gaussian focus,
z ~d 2 (Fig. 4a,b). Using light sculpting based on
TeFo (Oron et al. 2005, Zhu et al. 2005), this
issue can be addressed: The spectrum of a fem-
tosecond laser pulse is spatially dispersed by using
a grating. Imaging the spot on the grating onto
the sample with a telescope that consist of the
microscope objective and a lens, results in a con-
figuration where the spectral components of the
laser pulse overlap in time and space only within
a small axial region in which 2p excitation can oc-
cur. Thereby, axial sectioning can be achieved in-
dependently from the lateral size of the excitation
spot through dispersion and, thus, an effective
de-coupling of the axial and lateral confinement
of excitation can be realized (Fig. 4c).
In our
laboratory, we have used scanned TeFo (s-TeFo)
to shape the 3D PSF of our microscope near-
isotropically to match the required size to sam-
ple neuronal cell bodies in mouse cortex (~10-15
µm) (Prevedel et al. 2016). To generate sufficient
pulse energy for exciting the ~100 times larger
May 10, 2018
spot, a fiber amplified laser system was used. We
have shown that s-TeFo enabled unbiased single-
and dual-plane high-speed (up to 160 Hz) Ca2+
imaging as well as in vivo volumetric Ca2+ imag-
ing of a mouse cortical column (0.5 mm x 0.5 mm
x 0.5 mm) at single-cell resolution and fast volume
rates (3-6 Hz). An advantage of this approach is
that the peak intensities are lower because of the
larger spot size, which is beneficial for preventing
nonlinear tissue damage. On the other hand, the
larger spot size requires a higher average power, so
that s-TeFo operates closer to limits imposed by
heating, even though no detrimental effects due to
sample heating as reported through immunohisto-
chemistry were observed under the above condi-
tions. Finally, the reduced spatial resolution can
be compensated by using computational signal
extraction approaches to reduce possible mixing
of neuronal signals (Friedrich et al. 2017, Pnev-
matikakis et al. 2016).
5.1.2. 1D parallel acquisition:
A possibility to reduce the sequential scanning
to two dimensions is scanning a line along the
lateral or axial dimension instead of a focal spot
across the volume. One way is to generate a line,
which is scanned across the volume while the flu-
orescence signal is being recorded using an array
detector, e.g. a camera (Fig. 2d). In this case,
crosstalk due to scattering of the fluorescence light
from neighboring pixels in the parallel readout
limits the spatial resolution and the reachable
imaging depth. To improve this, the synchro-
nization of a line-shaped illumination with the
rolling shutter readout of modern CMOS cameras
can be exploited (Baumgart & Kubitscheck 2012,
Spiecker 2011). Our laboratory has demonstrated
the full potential of this method by combining the
line-scanning acquisition scheme with light sculpt-
ing using TeFo (Rupprecht et al. 2015). We could
demonstrate high-speed planar imaging of 200 x
200 µm FOV at 75 fps down to a depth of 75
µm in scattering brain tissue. By optimizing the
rolling shutter width, scattering effects could be
reduced by about a factor of three.
TeFo based excitation is part of the general cat-
egory of PSF engineering. Another instantiation
of PSF engineering are axially elongated PSFs.
They have been shown to enable simultaneous
recording along the axial direction and thereby re-
ducing the scanning to a 2D planar fashion (Fig.
2e). An example of such a technique uses Bessel
beams which are generated by an annular pattern
in the back-focal plane; the resulting interference
of plane waves at a shallow angle produces an axi-
ally elongated focus (Fig. 4d). Bessel beams have
been applied to 2p Ca2+ imaging in brain slices
(Botcherby et al. 2006, Thériault et al. 2014)
and recently also to volumetric in vivo functional
imaging (Lu et al. 2017). In the latter implemen-
tation, the neuronal sparsity in the brain volume
was exploited to ensure minimal structural over-
lap in the 2D projection. Neural activity could
be monitored at volume rates equivalent to the
2D frame rates, achieving up to 30 vol/s for a
~250 x 250 x 100 µm volume at depths down to
~160 µm in zebrafish larvae and fruit fly, as well
as scattering mouse and ferret cortex. Bessel foci
have a stronger lateral confinement than Gaussian
beams in the central peak for the same NA - but
the redistribution of power into the side rings re-
quires higher average power, in practice about 3-4
times more in the case of 2p excitation (Lu et al.
2017), and the fraction of power in the side rings
becomes larger with increasing NA. This is one of
the limitations of this approach as it leads to a
high fluorescence background when a one-photon
excitation strategy is used and to an increased
level of tissue heating in 2p excitation. Further-
more, both practical limitations in the achievable
aspect ratio of the focus and the increased mixing
of neuronal signals from different planes, as the
Bessel beam's axial length is increasing, restrict
the accessible axial range of this approach.
An interesting approach where a different type
of engineered PSF was used is vTwINS (Song et
al. 2017). Here, a volume was imaged using an
elongated, V-shaped PSF (Fig. 2f). Such a PSF
can be generated by impinging two small converg-
ing beams displaced from the optical axis onto
the back-focal plane of the microscope objective
(Fig. 4e). Due to the V-shaped focus, single neu-
rons appear as spatially separated pairs in the re-
sulting projection where the distance between the
May 10, 2018
projections encodes for the axial position in the
volume. Thereby, volumetric recordings from 550
x 550 x 50 µm at a speed of 30 vol/s could be
demonstrated. A disadvantage of this technique
is the increased background due to the V-shaped
PSF. Moreover, in this technique, the extent of
the axial range is limited as the success of de-
coding axial positions from the projection images
relies on sparsity, and thus restricts its application
to superficial volumes.
5.1.3. 2D parallel acquisition:
To further reduce scanning, 2D parallel acqui-
sition schemes record 2D images using a camera
while illuminating only a thin slice of the sample
volume orthogonally to the direction of observa-
tion (Fig. 2g) (Huisken et al. 2004, Voie et al.
1993). These techniques are generally referred to
as SPIM or 'light sheet' microscopy while there
are also other names for the different realizations
of this general idea (Santi 2011). SPIM comes in
various flavors that differ in the exact optical ar-
rangement, how the light sheet is generated and
how the signal is acquired. It has been used to
record the activity of the entire brain volume of
zebrafish larvae at ~1 Hz, capturing the activity
of the majority of all neurons at single-cell reso-
lution (Ahrens et al. 2013). The latest versions
of SPIM demonstrated Ca2+ imaging of the ze-
brafish brain with 420 x 830 x 160 µm FOV at 33
vol/s and cellular resolution which is more than an
order-of-magnitude improvement in volume rate
(Quirin et al. 2016). The light sheet is typically
generated using a cylindrical lens. While the lat-
eral resolution is determined by the wide-field de-
tection optics, the axial resolution is dictated by
the thickness of the light sheet. However, a thin,
high-NA light sheet diverges rapidly limiting the
FOV. An alternative is to generate the light sheet
using a scanning Bessel beam; 2 µm axial reso-
lution over a span of ~600 µm has been demon-
strated in the context of vasculature imaging in
zebrafish (Zhao et al. 2014). There are imple-
mentations of SPIM with both one-photon and
2p excitation; an advantage of SPIM with one-
photon excitation is that the required excitation
powers are much lower compared to 2p excitation.
However, one-photon excitation is limited to non-
or weakly scattering tissue because the scatter-
ing of the illumination beam strongly degrades
the illumination light sheet and impairs the spa-
tial resolution. The combination of 2p excitation
and SPIM allowed an increased SNR for imaging
in weakly scattering tissue (Truong et al. 2011,
Wolf et al. 2015). Another restriction is the fact
that SPIM requires optical access from at least
two orthogonal sides, in the case of opposing light
sheets even more (Lemon et al. 2015), to project
the excitation light sheet into the sample. The
dual-objective geometry limits biological applica-
tions to organisms such as zebrafish larvae where
optical access from different sites is possible, and
the organism needs to be immobilized making it
incompatible with behavioral studies.
Another technique that is based on the concept
of SPIM but requires optical access from only one
direction is swept confocally-aligned planar exci-
tation (SCAPE) (Bouchard et al. 2015, Kumar
et al. 2011). Here, an angled, swept light sheet is
used for excitation, and the detection is through
the same single microscope objective (Fig. 2h).
A confocal de-scanning and rotation mapping is
used to capture the scanned plane with a camera
achieving a volume rate of 10 vol/s for a FOV of
600 x 650 x 135 µm in scattering mouse brain. As
a one-photon modality, also SCAPE has a limited
penetration depth in scattering tissue.
One way to overcome the constrained sample
geometry of conventional LSM is to use 2p exci-
tation and light sculpting to introduce optical sec-
tioning for wide-field fluorescence imaging in epi-
configuration (Fig. 2i). Our laboratory demon-
strated wide-field TeFo where a disc-shaped ex-
citation area of ~75 µm diameter with an axial
confinement of ~2 µm was generated, and then
axially scanned (Schrödel et al. 2013). Using this
approach, we could for the first time demonstrate
whole-brain volumetric imaging in C. elegans by
recording from 75 x 75 x 40 µm at 13 vol/s. A
10 kHz-repetition rate laser system was used to
provide the required higher peak pulse energies of
~2 µJ at the sample for the wide-field excitation
(20 mW average power). Since the required aver-
age laser power scales linearly with the excitation
May 10, 2018
area, ultimately damage due to heating poses a
limitation to this technique.
5.1.4. 3D parallel acquisition:
A fully parallel acquisition modality captures
an entire volume in a single acquisition. This re-
quires the ability to map simultaneously all axial
information onto a single lateral plane that can
then be imaged in a parallel fashion.
One way to realize that experimentally is to im-
age different focal planes simultaneously using dif-
ferent cameras or different regions of a single cam-
era, which has been applied to functional imaging
in C. elegans using nine focal planes (Abrahams-
son et al. 2016) (Fig. 2j). In this one-photon ex-
citation realization, a volume of 40 x 40 x 18 µm
was imaged at 3 vol/s and diffraction-limited res-
olution. Custom diffractive optical elements are
required to split the focal planes and to correct for
spherical and chromatic aberrations. While this
approach does not require any computational re-
construction, the increasing complexity of the op-
tical system with more focal planes puts a practi-
cal limitation on the accessible volumes. A major
limitation of this type of multifocal microscopy is
the lack of optical sectioning; as a consequence,
both SNR and SBR are compromised by out-of-
focus fluorescence and because the fluorescence
signal is split over the number of focal planes.
Another technique for simultaneous 3D imag-
ing is LFM, which captures both the 2D location
and the 2D angular information of incident light
at the same time (Levoy et al. 2006, Lippmann
1908). A great advantage of LFM is that it is fully
scalable since the volume rate is fundamentally
independent of the volume size and only deter-
mined by the number of pixels and readout rate
of the camera. To realize LFM, a microlens array
is placed in the native image plane such that the
camera pixels can capture the rays of the light
field simultaneously (Fig. 2k). Thus, the PSF
encodes for the angular and, in consequence, the
axial information, and 3D volumes can be com-
putationally reconstructed using algorithms that
solve the inverse problem (Agard 1984, Broxton
et al. 2013). In our laboratory, we have shown the
application of LFM for whole-brain Ca2+ imaging
in small organisms like C. elegans and zebrafish
larvae (Prevedel et al. 2014). Recently, other vari-
ants of LFM have also been used to demonstrate
whole-brain Ca2+ imaging of freely swimming lar-
val zebrafish (Cong et al. 2017) and Drosophila
(Aimon et al. 2017).
4D phase-space measurements, as done by
LFM, contain a high redundancy in the infor-
mation on the 3D location of an object, re-
sulting in a strong robustness against scattering
compared to other camera-based wide-field tech-
niques (Liu et al. 2015). By using a compu-
tational approach termed Seeded Iterative De-
mixing (SID), our laboratory could extend the ca-
pabilities of LFM to mammalian cortex and show
large-volume functional recording in awake, be-
having mice (Nöbauer et al. 2017) reaching 30
vol/s at single cell resolution for volumes as large
as 900 x 900 x 260 µm and at depths down to
~380 µm.
It is noteworthy that the signal ex-
traction using SID considerably reduces crosstalk
between voxels compared to frame-by-frame im-
age reconstructions. Furthermore, SID reduces
the computational cost by about three orders of
magnitude.
An advantage of the one-photon excitation as
used in LFM is the low level of required average
power minimizing potential tissue damage. A lim-
itation of LFM is its reduced spatial resolution
as the microlens array trades lateral resolution
for the axial information, albeit potential signal
mixing can be addressed by computational ap-
proaches. The ultimate depth limit in LFM-based
methods is given by multiple scattering events
that will lead to a full loss of any directional infor-
mation of the emitted fluorescence photons and is
expected to be in the range of ~400-500 µm in the
cortex.
5.2. Sampling strategies: Unbiased versus
prior-based
All
imaging modalities discussed above are
based on an unbiased sampling strategy. This
means that they treat every voxel of the volume in
the same way – regardless if it contains a neuron
or not.
If prior knowledge on the specific neu-
ron locations is available, a deterministic, sparse
May 10, 2018
sampling strategy can be used, which will reduce
the required samples to the theoretical minimum
of one sample per neuron within the volume per
time unit. In cases where prior knowledge on only
average properties of the sample such as the spar-
sity is available, the number of spatial samples
can be reduced by a probabilistic approach.
One class of Ca2+ imaging that exploits spa-
tial sparseness and uses prior knowledge on neu-
ronal positions is RAM (Fernández-Alfonso et al.
2014, Katona et al. 2012, Reddy et al. 2008). In
this modality, fast scanners, such as AODs, are
used to access points in 3D that are known to
contain neurons or neuronal processes along ar-
bitrary trajectories in the volume (Fig. 2l). By
only recording at the a-priori known locations of
the neurons, acquisition rates up to ~50 points
per kHz can be achieved. A technical issue is that
AOD scanners have diffraction efficiencies of ~80-
90%. Since multiple (up to 4) AODs are used for a
full 3D RAM, this results in low power efficiency.
Moreover, AODs can only be used for small scan
angles, typically on the order of a few degrees,
limiting the FOV; and they also introduce strong
dispersion up to 72,000 fs2 (Katona et al. 2012).
Furthermore, the use of AODs puts limitations on
using multiple excitation wavelengths. A funda-
mental disadvantage of RAM is its higher suscep-
tibility to sample motion which makes its appli-
cation in awake, behaving animals difficult.
Another class of techniques using prior spa-
tial information are holographic techniques. They
typically generate multiple spots at arbitrary lo-
cations within the volume. In one realization, a
spatial light modulator in combination with tra-
ditional galvanometric scanners has been used to
simultaneously image multiple areas or alterna-
tively up to three axially separated layers within
the sample (Yang et al. 2016), Fig. 2m. This was
done by the spatial light modulator generating
multiple beamlets that are then scanned across
the FOV via the scanners. The fluorescence signal
is detected by a PMT which results in an image
that is effectively a super-position of individual
images stemming from the different beamlets. To
de-multiplex the signals, prior knowledge of the
neuron locations, the time-sparsity in their activ-
ity as well as a constraint non-negative matrix fac-
torization algorithm can be utilized. In practice,
the number of planes in this approach is limited
by the available laser power and, more fundamen-
tally, by the neuronal sparsity. Furthermore, the
separation between planes needs to be more than
~5 times the size of the axial PSF to avoid de-
creased optical sectioning and signal mixing.
Another holographic approach, called volume
projection imaging (VPI), uses an SLM to simul-
taneously excite neurons at locations which are
a-priori identified from a conventional 2p LSM
image stack (Quirin et al. 2014). Using a cus-
tom imaging system that employs a phase mask
in the imaging path, the PSF is modified such
that the image of each neuron becomes indepen-
dent of its axial position. Thus, the neurons can
be projected onto a single plane without out-of-
focus blur of the PSF and recorded on a cam-
era (Fig. 2n). This approach has been used to
demonstrate simultaneous 3D imaging within 280
x 270 x 110 µm at 30 vol/s in zebrafish. As with
all techniques relying on spatial priors, difficulties
arise with motion during the acquisition in awake,
behaving animals. Furthermore, prior knowledge
must be available, most likely by using a correla-
tive technique which might not always be possible.
5.3. Optical recording at greater depth
In the case that there is no or only weak scat-
tering, our above considerations regarding spatial
resolution and acquisition speed are universally
applicable to all sample locations.
In the pres-
ence of scattering and absorption this is no longer
true.
In general, the maximally obtainable imag-
ing depth z in optical microscopy is limited by
the scattering length l
in the medium, because
the excitation intensity is attenuated as I = I 0
exp(–z/l). In brain tissue, scattering is forward
directed with an anisotropy factor g ~0.9, result-
ing in photons that maintain some directional in-
formation even after a few scattering events (Ntzi-
achristos 2010). The scattering length in brain
tissue is ~50-100 µm in the visible part of the spec-
trum and ~100-400 µm in the NIR. Since available
fluorescent indicators have absorption and emis-
May 10, 2018
sion wavelengths in the visible part of the spec-
trum, single-photon excitation techniques are lim-
ited to a few hundred microns in practice; for cer-
tain NIR wavelengths a few millimeters can be
reached. In 2p microscopy, the visible indicators
can be excited with NIR light which can penetrate
deeper into scattering tissue. Secondly, the SBR
in 2p microscopy is significantly higher at greater
depths since 2p microscopy achieves a high local-
ization of excitation and an effective suppression
of out-of-focus fluorescence.
However, as fewer photons reach the focal vol-
ume with increasing tissue depth due to scatter-
ing, higher photon densities will be required at the
tissue surface to maintain the same level of SBR,
which will contribute to the background signal.
This will eventually limit the achievable depth in
2p microcopy. One way to achieve high SBR even
at greater tissue depths is to use 3p excitation.
Here, the higher degree of nonlinearity in the in-
teraction allows for a higher SBR compared to
2p microscopy since the SBR in 3p excitation is
given by SBR3p ∝ z 4NA6/(λ3l) exp(–z/l) com-
pared to SBR2p ∝ z 2NA2/(λ l) exp(–z/l) for 2p
excitation (Horton et al. 2013). 3p excitation is
also beneficial because the excitation wavelengths
for the available green and red Ca2+ indicators
lie within transparency windows of brain tissue
where absorption and scattering is reduced, allow-
ing to reach imaging depths beyond one millime-
ter. Recently, functional imaging of hippocampal
neurons at a depth of ~1 mm through the intact
mouse cortex has been demonstrated using 3p ex-
citation (Ouzounov et al. 2017). Since the 3p ex-
citation cross-section of fluorophores (~10−82 cm6
s2 / photon2) is considerably smaller than the 2p
equivalent (~10−49 cm4 s / photon), the 3p sig-
nal is more strongly affected by dispersion and
the associated pulse broadening as well as scatter-
induced reduction of photon density at the excita-
tion spot (Horton & Xu 2015). As a result, about
one order of magnitude higher pulse energies are
necessary in practice to produce a comparable sig-
nal level in 3p compared to 2p excitation which,
in turn, has to be considered in terms of tissue
heating (Wang et al. 2017).
An alternative approach is to alleviate the
underlying issue instead of avoiding it: When
enough information about the scattering process
is available, one can account for scattering us-
ing a corrected wave-front (Vellekoop & Mosk
2007). Wave front-shaping techniques accomplish
this by inverting the corresponding scattering ma-
trix (Conkey et al. 2012). This can be done by
using a fluorescent object of a known size, re-
ferred to as a guide star, and various types of
algorithms that infer the tissue-induced aberra-
tions from the recordings of the guide star and
use that information to shape the wave front of
the excitation beam such that it counters the ef-
fects of aberration at the focal plane. Different
methods using various types of algorithms and
guide stars have been demonstrated to implement
the general idea above (Horstmeyer et al. 2015).
This includes photo-acoustic techniques (Lai et al.
2015), the intrinsic 2p fluorescence signal (Katz et
al. 2014), or by focus scanning holographic aber-
ration probing (F-SHARP) (Papadopoulos et al.
2016). The required correction patterns for the
wave front are usually computed in an iterative
fashion and applied to the excitation beam via
amplitude or phase wave-front shaping using a
spatial light modulator (SLM) or a digital micro-
mirror device (DMD). The technical challenge in
these methods is that the wave-form correction is
only valid for a small FOV within the memory-
effect range of the scattering medium, which is
several tens of microns in brain tissue. Thus,
large FOVs would have to be tiled from individ-
ually corrected smaller FOVs. This means that a
large number of corrections have to be computed
for many sub-volumes, and they also need to be
applied fast enough while imaging. Furthermore,
in case of in vivo imaging, the correction is only
valid for a time period until the scattering chan-
nels have de-correlated due to tissue physiology
and vasculature or tissue deformation, for which
values on the order of 20 minutes have been re-
ported (Papadopoulos et al. 2016). When aberra-
tion correction methods are used in combination
with Ca2+ imaging, these effects either impact the
volume rate or the volumetric FOV at which neu-
ronal population recording can be practically re-
alized.
May 10, 2018
6. How to choose a microscopy method?
In the following section, we provide guidance
for navigating the main microscopy techniques
which have emerged over the recent years, and
for choosing the suitable method for a particu-
lar application. Assuming that the goal is to
achieve the highest possible neuronal sampling
rate for a given volume size, a maximally paral-
lel acquisition strategy - within the limits set by
photo-induced, compatible with the tissue scat-
tering properties - that is only as sequential as
necessary, should be used. Neurons or neuronal
processes should be spatially sampled as sparsely
as possible and as densely as necessary. If we as-
sume that it is desired to access the largest pos-
sible volume at the highest possible acquisition
speed, imaging modalities with a high degree of
parallelization should be chosen if the tissue ex-
hibits no or a low level of scattering. In the pres-
ence of scattering and depending on the desired
imaging depth, the methods of choice will be dom-
inated by techniques with an increasing level of
sequential acquisition.
Figure 5 provides a summary of the most widely
used microscopy techniques and their correspond-
ing areas of application. For instance, unbiased
imaging methods should be chosen for in vivo ap-
plications where significant sample motion is to
be expected such as in freely behaving or awake
head-fixed animals performing a motor tasks. In
contrast, when sample motion is limited or knowl-
edge about neuronal positions can be readily ob-
tained, prior-based sampling is preferred as it re-
duces the number of sampled voxels and thereby
increases the acquisition speed or the imaging vol-
ume. Another factor, as briefly discussed above,
is the degree of tissue scattering and recording
depth. In general, sequential methods, in partic-
ular 2p or multi-photon excitation, are preferred
in scattering tissue at the cost of reduced ac-
quisition speed. Methods using array detectors
such as cameras can be faster but are limited by
scattering and therefore more suited for non- or
weakly scattering samples. LFM-based methods
in combination with computational source extrac-
tion can considerably push the reachable depths
for camera-based modalities in scattering tissue
(Nöbauer et al. 2017).
7. Future developments
From the above discussions, it becomes clear
that there is a unique set of conditions and pa-
rameter space, within which each of the mi-
croscopy techniques operate optimally. Thus,
generally, hybrid methods can be expected to
cover a broader application space, since they have
the potential to combine advantages from multi-
ple complementary techniques. Moreover, there
are several frontiers for hardware improvement
which mutually contribute to the ability to image
larger volumes at faster acquisition rates. In the
following, we will briefly discuss recent progress
for faster scanners in LSM, cameras for parallel
recording modalities and laser technology.
A major limitation of the acquisition rates in
LSM methods is speed of the beam scanning.
When using mechanical scanning, the inertia is
usually the limiting factor. This becomes particu-
larly limiting in the case of axial scanning by mov-
ing the microscope objective. Thus, a strategy
when using mechanical scanning is to reduce in-
ertia by minimizing moving masses. For instance,
axial scanning of a heavy microscope objective
can be prevented by employing remote scanning
approaches (Botcherby et al.
2008), based on
scanning a light-weighted mirror via a voice coil
(Rupprecht et al. 2016, Sofroniew et al. 2016),
by using electro-tunable lenses (Chen et al. 2016,
Grewe et al. 2011), electro-wetting lenses, or ul-
trasound (TAG) lenses (Kong et al. 2015). One
way to increase the lateral scanning speed is to
replace the commonly used combination of a res-
onant galvanometer scanner, which can achieve
~16 kHz line scan rates, and a galvanometer
scanner with polygonal scanners (Bouchard et al.
2015, Li et al. 2017), or a combination of two res-
onant scanners that have a fixed phase relation
and can be used to scan on a Lissajous trajectory
(Newman et al. 2015). A variety of other tech-
nologies have been used for either lateral or axial
scanning, including AODs (Grewe et al. 2010),
electro-optic deflectors (Schneider et al. 2015),
May 10, 2018
Figure 5: Selection guide for Ca2+ imaging techniques for different sample conditions and biological applications. Meth-
ods are categorized regarding their requirement for scattering compatibility and suitability for awake imaging which
includes the susceptibility to motion artifacts. Example references for the techniques are provided.
and all-optical scanning (Wu et al. 2017). In the
case of parallel recording, ever-faster sensor ar-
rays with larger number of pixels, faster read-out
speed, and better light sensitivity are becoming
available. Progress in these technical areas will
further catalyze upscaling and the development
of novel optical techniques.
The possibility to record at increasing volume
rates over larger FOVs at sufficient SNR goes
hand in hand with progress in the illumination
sources. We believe that novel laser sources that
can provide high pulse energy with adjustable rep-
etition rates will be transformative as they allow
to flexibly adapt the light source to the design of
the imaging system for various applications and
sample conditions (Prevedel et al. 2016). Fiber
lasers or solid-state lasers at fixed wavelengths
optimized for the wavelengths of well-engineered
fluorophores will prove to be more useful than
tunable titanium sapphire lasers. Lasers based
on phenomena such as soliton self-frequency shift
(Gruner-Nielsen et al. 2010, Zhu et al. 2013)
are starting to become commercially available and
will give access to a broader range of more wave-
lengths – especially in the infrared. Ultimately,
all high-speed Ca2+ imaging techniques will chal-
lenge fundamental limits such as energy dissipa-
tion (Marblestone et al.
2013) and constrains
given by the properties of the currently used flu-
orescent probes. Progress in imaging techniques
will also incentivize further developments in the
probe development. We expect that together they
will further push the boundaries of current pos-
sibilities for manipulating and recoding neuronal
network activity from large population at high
speed and resolution.
Acknowledgements
We would like to acknowledge support from
the National Institute of Neurological Disorders
and Stroke of the US National Institutes of
Health under Award Number U01NS103488 and
U01NS094263-01 as well as The Kavli Founda-
tion. S.W. has been supported by a Leon Levy
Fellowship at The Rockefeller University.
Posted with permission from the Annual Re-
view of Neuroscience, Volume 41 ©2018 by An-
nual Reviews, http://www.annualreviews.org/.
List of abbreviations
2p: two-photon
3p: three-photon
AOD: acousto-optical deflector
Ca2+: calcium
fMRI: functional magnetic resonance imaging
May 10, 2018
FOV: field of view
GECI: Genetically encoded calcium indicator
GFP: green fluorescent protein
LFM: light-field microscopy
LSM: laser scanning microscopy
NA: numerical aperture
PSF: point spread function
RAM: random access microscopy
SNR: signal-to-noise ratio
SBR: signal-to-background ratio
SPIM: selective plane illumination microscopy
TeFo: temporal focusing
References
Abrahamsson S, Ilic R, Wisniewski J, Mehl B,
Yu L, et al. 2016. Multifocus microscopy with
precise color multi-phase diffractive optics applied
in functional neuronal imaging. Biomed. Opt.
Express. 7(3):855–15
Agard D. 1984. Optical Sectioning Microscopy:
Cellular Architecture in Three Dimensions. An-
nual Review of Biophysics and Biomolecular
Structure. 13(1):191–219
Ahrens MB, Orger MB, Robson DN, Li JM,
Keller PJ. 2013. Whole-brain functional imaging
at cellular resolution using light-sheet microscopy.
Nat Meth. 10(5):413–20
Aimon S, Katsuki T, Grosenick L, Broxton M,
Deisseroth K, et al. 2017. Fast whole brain imag-
ing in adult Drosophila during response to stimuli
and behavior. bioRxiv. 1–34
Akerboom J, Chen TW, Wardill TJ, Tian
L, Marvin JS, et al.
Optimiza-
tion of a GCaMP Calcium Indicator for Neu-
ral Activity Imaging. Journal of Neuroscience.
32(40):13819–40
2012.
Amir W, Carriles R, Hoover EE, Planchon TA,
Durfee CG, Squier JA. 2007. Simultaneous imag-
ing of multiple focal planes using a two-photon
scanning microscope. Opt. Lett. 32(12):1731–34
Averbeck BB, Latham PE, Pouget A. 2006.
Neural correlations, population coding and com-
putation. Nat Rev Neurosci. 7(5):358–66
Baumgart E, Kubitscheck U. 2012. Scanned
light sheet microscopy with confocal slit detec-
tion. Opt. Express. 20(19):21805–10
Berenyi A, Somogyvari Z, Nagy AJ, Roux L,
Long JD, et al. 2014. Large-scale, high-density
(up to 512 channels) recording of local circuits
in behaving animals. Journal of Neurophysiology.
111(5):1132–49
Botcherby EJ, Juškaitis R, Booth MJ, Wilson
T. 2008. An optical technique for remote fo-
cusing in microscopy. Optics Communications.
281(4):880–87
Botcherby EJ, Juškaitis R, Wilson T. 2006.
Scanning two photon fluorescence microscopy
with extended depth of field. Optics Communica-
tions. 268(2):253–60
Bouchard MB, Voleti V, Mendes CS, Lacefield
C, Grueber WB, et al. 2015. Swept confocally-
aligned planar excitation (SCAPE) microscopy
for high-speed volumetric imaging of behaving or-
ganisms. Nature Photonics. 9(2):113–19
Boyden ES, Zhang F, Bamberg E, Nagel G,
Deisseroth K. 2005. Millisecond-timescale, genet-
ically targeted optical control of neural activity.
Nat Neurosci. 8(9):1263–68
Braitenberg V, Schüz A. 1998. Cortex: Statis-
tics and Geometry of Neuronal Connectivity.
Berlin, Heidelberg: Springer Berlin Heidelberg
Broxton M, Grosenick L, Yang S, Cohen N, An-
dalman A, et al. 2013. Wave optics theory and
3-D deconvolution for the light field microscope.
Opt. Express. 21(21):25418–22
Buzsaki G. 2004. Large-scale recording of neu-
ronal ensembles. Nat Neurosci. 7(5):446–51
Buzsaki G, Mizuseki K. 2014.
The log-
dynamic brain: how skewed distributions affect
network operations. Nature Publishing Group.
15(4):264–78
Chamberland S, Yang HH, Pan MM, Evans
SW, Guan S, et al.
2017. Fast two-photon
imaging of subcellular voltage dynamics in neu-
ronal tissue with genetically encoded indicators.
eLIFE. 1–76
Chan KY, Jang MJ, Yoo BB, Greenbaum A,
Ravi N, et al. 2017. Engineered AAVs for ef-
ficient noninvasive gene delivery to the central
and peripheral nervous systems. Nat Neurosci.
20(8):1172–79
Chen JL, Voigt FF, Javadzadeh M, Krueppel
R, Helmchen F. 2016. Long-range population dy-
May 10, 2018
namics of anatomically defined neocortical net-
works. eLIFE. 1–26
Chen T-W, Wardill TJ, Sun Y, Pulver SR, Ren-
ninger SL, et al. 2013. Ultrasensitive fluorescent
proteins for imaging neuronal activity. Nature.
499(7458):295–300
Cheng A, Gonçalves JT, Golshani P, Arisaka
K, Portera-Cailliau C. 2011.
Simultaneous
two-photon calcium imaging at different depths
with spatiotemporal multiplexing. Nat Meth.
8(2):139–42
Cong L, Wang Z, Chai Y, Hang W, Shang C,
et al. 2017. Rapid whole brain imaging of neural
activity in freely behaving larval zebrafish (Danio
rerio). 1–55
Conkey DB, Caravaca-Aguirre AM, Piestun R.
2012. High-speed scattering medium characteri-
zation with application to focusing light through
turbid media. Opt. Express. 20(2):1733–38
Denk W, Strickler JH, Webb WW. 1990. Two-
Photon Laser Scanning Fluorescence Microscopy.
Science. 1–5
Dittgen T, Nimmerjahn A, Komai S, Licznerski
P, Waters J, et al. 2004. Lentivirus-based genetic
manipulations of cortical neurons and their op-
tical and electrophysiological monitoring in vivo.
Proceedings of the National Academy of Sciences.
101(52):18206–11
Fernández-Alfonso T, Nadella KMNS, Iacaruso
MF, Pichler B, Roš H, et al. 2014. Monitor-
ing synaptic and neuronal activity in 3D with
synthetic and genetic indicators using a compact
acousto-optic lens two-photon microscope. Jour-
nal of Neuroscience Methods. 222:69–81
Foust KD, Nurre E, Montgomery CL, Hernan-
dez A, Chan CM, Kaspar BK. 2008.
Intravas-
cular AAV9 preferentially targets neonatal neu-
rons and adult astrocytes. Nature Biotechnology.
27(1):59–65
Friedrich J, Yang W, Soudry D, Mu Y, Ahrens
MB, et al. 2017. Multi-scale approaches for high-
speed imaging and analysis of large neural popu-
lations. PLoS Comput Biol. 13(8):e1005685–24
Gouwens NW, Wilson RI. 2009. Signal Propa-
gation in Drosophila Central Neurons. Journal of
Neuroscience. 29(19):6239–49
Grewe BF, Langer D, Kasper H, Kampa BM,
Helmchen F. 2010.
High-speed in vivo cal-
cium imaging reveals neuronal network activ-
ity with near-millisecond precision. Nat Meth.
7(5):399–405
Grewe BF, Voigt FF, van Hoff M, Helmchen F.
2011. Fast two-layer two-photon imaging of neu-
ronal cell populations using an electrically tunable
lens. Biomed. Opt. Express. 1–12
Grienberger C, Konnerth A. 2012.
Imaging
Calcium in Neurons. Neuron. 73(5):862–85
Grinvald A, Salzberg BM, Lev-Ram V,
Hildesheim R. 1987. Optical recording of synap-
tic potentials from processes of single neurons us-
ing intracellular potentiometric dyes. Biophysj.
51(4):643–51
Gruner-Nielsen L, Jakobsen D, Jespersen KG,
Pálsdóttir B. 2010. A stretcher fiber for use in
fs chirped pulse Yb amplifiers. Opt. Express.
18(4):3768–6
Grynkiewicz G, Poenie M, Tsien RY. 1985.
New Generation of Ca2+ Indicators with Greatly
Improved Fluorescence Properties. The Journal
of Biological Chemistry. 3440
Harris KD, Quiroga RQ, Freeman J, Smith SL.
2016. Improving data quality in neuronal popu-
lation recordings. Nat Neurosci. 19(9):1165–74
Herculano-Houzel S, Collins CE, Wong P, Kaas
JH. 2007. Cellular scaling rules for primate
brains. Proceedings of the National Academy of
Sciences. 104(9):3562–67
Herculano-Houzel S, Mota B, Lent R. 2006.
Cellular scaling rules for rodent brains. Pro-
ceedings of the National Academy of Sciences.
103(32):12138–43
Hilgen G, Sorbaro M, Pirmoradian S, Muth-
mann J-O, Kepiro IE, et al. 2017. Unsupervised
Spike Sorting for Large-Scale, High- Density Mul-
tielectrode Arrays. CellReports. 18(10):2521–32
Hochbaum DR, Zhao Y, Farhi SL, Klapoetke N,
Werley CA, et al. 2014. All-optical electrophysi-
ology in mammalian neurons using engineered mi-
crobial rhodopsins. Nat Meth. 11(8):825–33
Hopt A, Neher E. 2001. Highly Nonlinear
Photodamage in Two-Photon Fluorescence Mi-
croscopy. Biophysical Journal. 1–8
Horstmeyer R, Ruan H, Yang C. 2015.
Guidestar-assisted wavefront-shaping methods for
May 10, 2018
focusing light into biological tissue. Nature Pho-
tonics. 1–9
Horton NG, Wang K, Kobat D, Clark CG, Wise
FW, et al. 2013. SI: In vivo three-photon mi-
croscopy of subcortical structures within an intact
mouse brain. Nature Photonics. 7(3):205–9
Horton NG, Xu C. 2015. Dispersion compen-
sation in three-photon fluorescence microscopy at
1,700 nm. Biomed. Opt. Express. 6(4):1392–96
Howarth C, Gleeson P, Attwell D. 2012. Up-
dated energy budgets for neural computation in
the neocortex and cerebellum. 32(7):1222–32
Huettel SA, Song AW, McCarthy GJ. 2008.
Functional Magnetic Resonance Imaging. Sinauer
Associates. Second Edition ed.
Huisken J, Swoger J, Del Bene F, Wittbrodt J,
Stelzer EHK. 2004. Optical Sectioning Deep In-
side Live Embryos by Selective Plane Illumination
Microscopy. Science. 305:1–18
Insel TR, Landis SC, Collins FS. 2013. The
NIH BRAIN Initiative. Science. 340:687
Ji G, Feldman ME, Deng K-Y, Greene KS, Wil-
son J, et al. 2004. Ca 2+-sensing Transgenic
Mice.
JOURNAL OF BIOLOGICAL CHEM-
ISTRY. 279(20):21461–68
Kalmbach AS, Waters J. 2012. Brain surface
temperature under a craniotomy. Journal of Neu-
rophysiology. 108(11):3138–46
Kato S, Kaplan HS, Schrödel T, Skora S, Lind-
say TH, et al. 2015. Global Brain Dynamics Em-
bed the Motor Command Sequence of Caenorhab-
ditis elegans. Cell. 163(3):656–69
Katona G, Szalay G, Maák P, Kaszas A, Veress
M, et al. 2012. Fast two-photon in vivo imaging
with three-dimensional random-access scanning
in large tissue volumes. Nat Meth. 9(2):201–8
Katz O, Small E, Guan Y, Silberberg Y.
2014. Noninvasive nonlinear focusing and imaging
through strongly scattering turbid layers. Optica.
1(3):170–75
Kerr JND, Denk W. 2008.
Imaging in vivo:
watching the brain in action. Nat Rev Neurosci.
9(3):195–205
Koester HJ, Baur D, Uhl R, Hell SW. 1999.
Fluorescence Imaging with Pico- and Femtosec-
ond Two-Photon Excitation: Signal and Photo-
damage. Biophysical Journal. 1–11
Kong L, Tang J, Little JP, Yu Y, Lämmermann
T, et al. 2015. Continuous volumetric imaging
via an optical phase-locked ultrasound lens. Nat
Meth. 12(8):759–62
Kovacevic N. 2004. A Three-dimensional MRI
Atlas of the Mouse Brain with Estimates of
the Average and Variability. Cerebral Cortex.
15(5):639–45
Krüger J. 2005.
Simultaneous individual
recordings from many cerebral neurons: Tech-
niques and results. In Reviews of Physiology, Bio-
chemistry and Pharmacology, Volume 98, Vol. 98,
pp. 177–233. Berlin, Heidelberg: Springer Berlin
Heidelberg
Kumar S, Wilding D, Sikkel MB, Lyon AR,
MacLeod KT, Dunsby C. 2011. High-speed 2D
and 3D fluorescence microscopy of cardiac my-
ocytes. Opt. Express. 19(15):13839–39
Lai P, Wang L, Tay JW, Wang LV. 2015. Pho-
toacoustically guided wavefront shaping for en-
hanced optical focusing in scattering media. Na-
ture Photonics. 9(2):126–32
Lemon WC, ckendorf BHO, McDole K, Bran-
son K, Freeman J, et al.
2015. Whole-
central nervous system functional imaging in lar-
val Drosophila. Nature Communications. 6:1–16
Levoy M, Ng R, Adams A, Footer M, Horowitz
M. 2006. Light Field Microscopy. AMC. 1–11
Li YX, Gautam V, Brüstle A, Cockburn IA,
Daria VR, et al. 2017. Flexible polygon-mirror
based laser scanning microscope platform for mul-
tiphoton in-vivoimaging. J. Biophoton. 2:143–12
Lin MZ, Schnitzer MJ. 2016. Genetically en-
coded indicators of neuronal activity. Nat Neu-
rosci. 19(9):1142–53
Lippmann MG. 1908. Épreuves réversibles.
Computes Rendos
Photographies intégrales.
delAcademie des Sciences. 146:446–51
Liu H-Y, Jonas E, Tian L, Zhong J, Recht B,
Waller L. 2015. 3D imaging in volumetric scatter-
ing media using phase-space measurements. Opt.
Express. 23(11):14461–11
Lu R, Sun W, Liang Y, Kerlin A, Bierfeld J, et
al. 2017. Video-rate volumetric functional imag-
ing of the brain at synaptic resolution. Nat Neu-
rosci
Mao T, O'Connor DH, Scheuss V, Nakai J, Svo-
May 10, 2018
boda K. 2008. Characterization and Subcellular
Targeting of GCaMP-Type Genetically-Encoded
Calcium Indicators. PLoS ONE. 3(3):e1796–10
Marblestone AH, Zamft BM, Maguire YG,
Shapiro MG, Cybulski TR, et al. 2013. Physical
principles for scalable neural recording. Frontiers
in Computational Neuroscience. 7:
Martin, Jonathan D Thiessen, Laryssa M Kur-
jewicz, Shelley L Germscheid, Allan J Turner, et
al. 2010. Longitudinal Brain Size Measurements
in APP/PS1 Transgenic Mice. MRI. 19–8
Miyawaki A, Llopis J, Heim R, McCaffery JM,
Adams JA, et al. 1997. Fluorescent indicators
for Ca2+based on green fluorescent proteins and
calmodulin. Nature. 388(6645):882–87
Mukamel EA, Nimmerjahn A, Schnitzer MJ.
2009. Automated Analysis of Cellular Signals
from Large-Scale Calcium Imaging Data. Neu-
ron. 63(6):747–60
Nagai T, Yamada S, Tominaga T, Ichikawa M,
Miyawaki A. 2004. Expanded dynamic range of
fluorescent indicators for Ca2_ by circularly per-
muted yellow fluorescent proteins. Proceedings of
the National Academy of Sciences. 1–6
Nagel G, Ollig D, Fuhrmann M, Kateriya S,
Musti AM, et al. 2002. Channelrhodopsin-1: A
Light-Gated Proton Channel in Green Algae. Sci-
ence. 296(5577):2395–98
Nakai J, Ohkura M, Imoto K. 2001. A high
signal-to-noise Ca2+ probe composed of a single
green fluorescent protein. Nature Biotechnology.
19:137
Naumann EA, Kampff AR, Prober DA, Schier
AF, Engert F. 2010. Monitoring neural activ-
ity with bioluminescence during natural behavior.
Nature Publishing Group. 13(4):513–20
Newman JA, Sullivan SZ, Muir RD, Sreehari S,
Bouman CA, Simpson GJ. 2015. Multi-channel
beam-scanning imaging at kHz frame rates by
Lissajous trajectory microscopy.
SPIE BiOS.
9330:933009–8
Nöbauer T, Skocek O, Pernía-Andrade AJ,
Weilguny L, Traub FM, et al.
2017. Video
rate volumetric Ca2+ imaging across cortex us-
ing seeded iterative demixing (SID) microscopy.
Nature Publishing Group. 1–10
Ntziachristos V. 2010. Going deeper than mi-
croscopy: the optical imaging frontier in biology.
Nature Publishing Group. 7(8):603–14
Ogawa S, Lee TM, Kay AR, Tank DW. 1990.
Brain magnetic resonance imaging with con-
trast dependent on blood oxygenation.
Pro-
ceedings of the National Academy of Sciences.
87(24):9868–72
Oron D, Tal E, Silberberg Y. 2005. Scanning-
less depth-resolved microscopy. Opt. Express.
1–9
Ouzounov DG, Wang T, Wang M, Feng DD,
Horton NG, et al. 2017.
In vivo three-photon
imaging of activity of GCaMP6-labeled neurons
deep in intact mouse brain. Nature Publishing
Group. 1–5
Papadopoulos IN, Jouhanneau J-S, Poulet JFA,
Judkewitz B. 2016. Scattering compensation by
focus scanning holographic aberration probing (F-
SHARP). Nature Photonics. 1–9
Pawley JB. 2006. Handbook of Biological Con-
focal Microscopy. Springer. 3rd ed.
Pnevmatikakis EA, Soudry D, Gao Y, Machado
TA, Merel J, et al. 2016. Simultaneous Denoising,
Deconvolution, and Demixing of Calcium Imaging
Data. Neuron. 89(2):285–99
Podgorski K, Ranganathan GN. 2016. Brain
heating induced by near infrared lasers during
multi-photon microscopy. Journal of Neurophys-
iology. jn.00275.2016–12
Prevedel R, Verhoef AJ, Weisenburger S,
2016.
Pernía-Andrade AJ, Huang BS, et al.
Fast volumetric calcium imaging across multiple
cortical layers using sculpted light. Nat Meth.
13(12):1021–28
Prevedel R, Yoon Y-G, Hoffmann M, Pak N,
Wetzstein G, et al. 2014. Simultaneous whole-
animal 3D imaging of neuronal activity using
light-field microscopy. Nat Meth. 11(7):727–30
Quirin S, Jackson J, Peterka DS, Yuste R.
2014. Simultaneous imaging of neural activity
in three dimensions. Front. Neural. Circuits.
8(65):413–11
Quirin S, Vladimirov N, Yang C-T, Peterka DS,
Yuste R, B Ahrens M. 2016. Calcium imaging of
neural circuits with extended depth-of-field light-
sheet microscopy. Opt. Lett. 41(5):855–4
Reddy GD, Kelleher K, Fink R, Saggau P.
May 10, 2018
2008. Three-dimensional random access multi-
photon microscopy for functional imaging of neu-
ronal activity. Nat Neurosci. 11(6):713–20
Rein K, Zöckler M, Mader MT, Grübel C,
Heisenberg M. 2002. The Drosophila Standard
Brain. Current Biology. 12(3):227–31
Rey HG, Pedreira C, Quiroga RQ. 2015. Past,
present and future of spike sorting techniques.
Brain Research Bulletin. 119(Part B):106–17
Rupprecht P, Prendergast A, Wyart C,
Friedrich RW. 2016. Remote z-scanning with a
macroscopic voice coil motor for fast 3D multi-
photon laser scanning microscopy. Biomed. Opt.
Express. 7(5):1656–16
Rupprecht P, Prevedel R, Groessl F, Hauben-
sak WE, Vaziri A. 2015. Optimizing and extend-
ing light-sculpting microscopy for fast functional
imaging in neuroscience. Biomed. Opt. Express.
6(2):353–16
Sahin B, Aslan H, Unal B, Canan S, Bilgic S, et
al. 2001. Brain volumes of the lamb, rat and bird
do not show hemispheric asymmetry: a stereolog-
ical study. Image Analysis Stereology. 20(1):1–5
Salzberg BM, Grinvald A, Cohen LB, Davila
HV, Ross WN. 1977. Optical Recording of Neu-
ronal Activity in an Invertebrate Central Nervous
System: Simultaneous Monitoring of Several Neu-
rons. Journal of Neurophysiology
Santi PA. 2011. Light Sheet Fluorescence Mi-
croscopy. J Histochem Cytochem. 59(2):129–38
Schneider J, Zahn J, Maglione M, Sigrist SJ,
Marquard J, et al. 2015. Ultrafast, temporally
stochastic STED nanoscopy of millisecond dy-
namics. Nat Meth. 12(9):827–30
Schrödel T, Prevedel R, Aumayr K, Zimmer
M, Vaziri A. 2013. Brain-wide 3D imaging of
neuronal activity in Caenorhabditis elegans with
sculpted light. Nat Meth. 10(10):1013–20
Shimogori T, Ogawa M. 2008. Gene application
with in uteroelectroporation in mouse embryonic
brain. Development, Growth & Differentiation.
50(6):499–506
Shimomura O, Johnson FH, Saiga Y. 1962. Ex-
traction, Purification and Properties of Aequorin,
a Bioluminescent Protein from the Luminous Hy-
dromedusan,Aequorea. J. Cell. Comp. Physiol.
59(3):223–39
Simpson JH. 2009. Mapping and Manipulating
Neural Circuits in the Fly Brain. In Genetic Dis-
section of Neural Circuits and Behavior, Vol. 65,
pp. 79–143. Elsevier Inc. 1st ed.
So P, Dong CY, Masters B, Berland K. 2000.
Two-photon excitation fluorescence microscopy.
Annu. Rev. Neurosci.
Sofroniew NJ, Flickinger D, King J, Svoboda K.
2016. A large field of view two-photon mesoscope
with subcellular resolution for in vivo imaging.
eLIFE. 1–20
Song A, Charles AS, Koay SA, Gauthier JL,
Thiberge SY, et al. 2017. Volumetric two-photon
imaging of neurons using stereoscopy (vTwINS).
Nature Publishing Group. 14(4):420–26
Spiecker H. 2011. Verfahren und Anordnung
Trebino R, Zeek E. 2000. Ultrashort Laser
Pulses.
In Frequency-Resolved Optical Gating:
the Measurement of Ultrashort Laser Pulses, pp.
11–35. Boston, MA: Springer US
Truong TV, Supatto W, Koos DS, Choi JM,
Fraser SE. 2011. Deep and fast live imaging with
two-photon scanned light-sheet microscopy. Nat
Meth. 8(9):757–60
Vaziri A, Emiliani V. 2012. Reshaping the op-
May 10, 2018
zur Mikroskopie. DE102010013223A1
Stirman JN, Smith IT, Kudenov MW, Smith
SL. 2016. Wide field-of-view, multi-region,
two-photon imaging of neuronal activity in
the mammalian brain. Nature Biotechnology.
34(8):865–70
Tallini YN, Ohkura M, Choi BR, Ji G, Imoto K,
et al. 2006. Imaging cellular signals in the heart
in vivo: Cardiac expression of the high-signal
Ca2+ indicator GCaMP2. Proceedings of the Na-
tional Academy of Sciences. 103(12):4753–58
Thériault G, Cottet M, Castonguay A, Mc-
Carthy N, De Koninck Y. 2014. Extended two-
photon microscopy in live samples with Bessel
beams: steadier focus, faster volume scans, and
simpler stereoscopic imaging. Frontiers in Cellu-
lar Neuroscience. 1–11
Tian L, Hires SA, Mao T, Huber D, Chiappe
ME, et al.
Imaging neural activity in
worms, flies and mice with improved GCaMP
calcium indicators. Nature Publishing Group.
6(12):875–81
2009.
tical dimension in optogenetics. Current Opinion
in Neurobiology. 22(1):128–37
Vellekoop IM, Mosk AP. 2007. Focusing co-
herent light through opaque strongly scattering
media. Opt. Lett. 32(16):2309–11
Voie AH, Burns DH, Spelman FA. 1993.
section-
Orthogonal-plane fluorescence optical
ing: Three-dimensional imaging of macroscopic
biological specimens.
Journal of Microscopy.
170(3):229–36
Wang T, Ouzounov D, Wang M, Xu C.
2017. Quantitative Comparison of Two-photon
and Three-photon Activity Imaging of GCaMP6s-
labeled Neurons in vivo in the Mouse Brain.
Brain, p. BrM4B.4. Washington, D.C.: OSA
Weisenburger S, Sandoghdar V. 2015. Light
microscopy: an ongoing contemporary revolution.
Contemporary Physics. 56(2):123–43
White JG, Southgate E, Thomson JN, Brenner
S. 1986. The Structure of the Nervous System
of the Nematode Caenorhabditis elegans. Phil.
Trans. R. Soc. B. 314(1165):1–340
Wolf S, Supatto W, Debrégeas G, Mahou P,
Kruglik SG, et al. 2015. Whole-brain functional
imaging with two-photon light-sheet microscopy.
Nature Publishing Group. 12(5):379–80
Wu J, Tang AHL, Mok ATY, Yan W, Chan
GCF, et al.
2017. Multi-MHz laser-scanning
single-cell fluorescence microscopy by spatiotem-
porally encoded virtual source array. Biomed.
Opt. Express. 8(9):4160–12
Xu Y, Zou P, Cohen AE. 2017. Voltage imag-
ing with genetically encoded indicators. Current
Opinion in Chemical Biology. 39:1–10
Yang W, Miller J-EK, Carrillo-Reid L, Pnev-
matikakis E, Paninski L, et al. 2016. Simultane-
ous Multi-plane Imaging of Neural Circuits. Neu-
ron. 89(2):269–84
Yuste R, Bargmann C. 2017. Toward a Global
BRAIN Initiative. Cell. 168(6):956–59
Zariwala HA, Borghuis BG, Hoogland TM,
Madisen L, Tian L, et al. 2012. A Cre-Dependent
GCaMP3 Reporter Mouse for Neuronal Imaging
In Vivo. Journal of Neuroscience. 32(9):3131–41
Zhang F, Aravanis AM, Adamantidis A, de
Lecea L, Deisseroth K. 2007. Circuit-breakers:
optical technologies for probing neural signals and
systems. Nat Rev Neurosci. 8(9):732–32
Zhao M, Zhang H, Li Y, Ashok A, Liang R,
et al. 2014. Cellular imaging of deep organ us-
ing two-photon Bessel light-sheet nonlinear struc-
tured illumination microscopy. Biomed. Opt. Ex-
press. 5(5):1296–13
Zhu G, van Howe J, Durst M, Zipfel W, Xu C.
2005. Simultaneous spatial and temporal focusing
of femtosecond pulses. Opt. Express. 1–7
Zhu L, Verhoef AJ, Jespersen KG, Kalashnikov
VL, Grüner-Nielsen L, et al. 2013. Generation of
high fidelity 62-fs, 7-nJ pulses at 1035 nm from a
net normal-dispersion Yb-fiber laser with anoma-
lous dispersion higher-order-mode fiber. Opt. Ex-
press. 21(14):16255–58
May 10, 2018
|
1904.01360 | 1 | 1904 | 2019-04-02T12:11:32 | Melting transitions in biomembranes | [
"physics.bio-ph"
] | We investigated melting transitions in biological membranes in their native state that include their membrane proteins. These membranes originated from \textit{E. coli}, \textit{B. subtilis}, lung surfactant and nerve tissue from the spinal cord of several mammals. For some preparations, we studied the pressure, pH and ionic strength dependence of the transition. For porcine spine, we compared the transition of the native membrane to that of the extracted lipids. All preparations displayed melting transitions of 10-20 degrees below physiological or growth temperature, independent of the organism of origin and the respective cell type. The position of transitions in \textit{E. coli} membranes depends on the growth temperature. We discuss these findings in the context of the thermodynamic theory of membrane fluctuations that leads to largely altered elastic constants, an increase in fluctuation lifetime and in membrane permeability associated with the transitions. We also discuss how to distinguish lipid transitions from protein unfolding transitions. Since the feature of a transition slightly below physiological temperature is conserved even when growth conditions change, we conclude that the transitions are likely to be of major biological importance for the survival and the function of the cell. | physics.bio-ph | physics |
Melting transitions in biomembranes
Tea Muzi´c, Fatma Tounsi, Søren B. Madsen, Denis Pollakowski, Manfred Konrad and Thomas Heim-
burg
Niels Bohr Institute, University of Copenhagen, Blegdamsvej 17, 2100 Copenhagen Ø, Denmark
ABSTRACT We investigated melting transitions in biological membranes in their native state that include their membrane
proteins. These membranes originated from E. coli, B. subtilis, lung surfactant and nerve tissue from the spinal cord of several
mammals. For some preparations, we studied the pressure, pH and ionic strength dependence of the transition. For porcine
spine, we compared the transition of the native membrane to that of the extracted lipids. All preparations displayed melting
transitions of 10-20 degrees below physiological or growth temperature, independent of the organism of origin and the
respective cell type. The position of transitions in E. coli membranes depends on the growth temperature. We discuss these
findings in the context of the thermodynamic theory of membrane fluctuations that leads to largely altered elastic constants,
an increase in fluctuation lifetime and in membrane permeability associated with the transitions. We also discuss how to
distinguish lipid transitions from protein unfolding transitions. Since the feature of a transition slightly below physiological
temperature is conserved even when growth conditions change, we conclude that the transitions are likely to be of major
biological importance for the survival and the function of the cell.
Keywords: thermodynamics; fluctuations; E.coli; lung surfactant; B.subtilis; nerves; elastic constants
∗corresponding author, [email protected].
Introduction
Lipid bilayers display melting transitions at a temperature
Tm, during which both lateral and chain order change ac-
companied by the absorption of heat. The typical range of
transition temperatures is from -20◦ to +60◦C depending on
chain length, chain saturation and the chemical nature of the
head groups (1). Melting transitions can easily be observed
with calorimeters and various spectroscopic methods such as
infrared spectroscopy or magnetic resonance. Mixtures of
lipids with different melting temperatures display phase be-
havior that can be plotted in phase diagrams (2). These di-
agrams show the coexistence of phases as a function of both
molar fractions of the components and intensive variables such
as temperature and pressure. From the theoretical analysis of
phase diagrams one can obtain melting profiles of the lipid
mixtures, and understand the phase separation processes (1,
3). Consequently, one finds domain formation in certain re-
gimes of the phase diagram (4, 5).
It is little known that biological membranes also display
melting transitions close to the physiological temperature re-
gime. Several publications in the 1970s reported, for exam-
ple, melting phenomena in the membranes of Mycoplasma
laidlawii (6, 7), Micrococcus lysodeikticus (8), mouse fibrob-
last LM cells (9), and red blood cells (10). Haest et al. (11)
showed by electron microscopy that the arrangement of pro-
teins in a native bacterial membrane correlates with this tran-
sition. Transitions have also been reported for lung surfactant
(12) and E.coli membranes (13). However, the relevance of
these transitions for the function of cells has been little ap-
preciated. Since biological membranes consist of hundreds
or even thousands of components (14, 15), it seems impossi-
ble to construct phase diagrams. Due to Gibbs phase rule, the
possible number of coexisting phases is of similar order as
the number of components when neglecting the phase bound-
aries. The compositions of domains cannot easily be derived
from simple physical considerations or measurements if fi-
nite size domains are considered. We show below that it is
nevertheless possible to extract useful information from the
melting profiles.
1
Melting transitions strongly influence elastic constants
(16, 17). According to the fluctuation-dissipation theorem,
the heat capacity is proportional to enthalpy fluctuations. All
other susceptibilities are similarly related to fluctuations of
extensive variables. For instance, the isothermal area com-
pressibility is proportional to area fluctuations (16, 18) and
the capacitance is proportional to charge fluctuations (19).
Further, the isothermal compressibility is proportional to the
heat capacity (16, 17, 20). The same is true for the other sus-
ceptibilities. They are all related to the heat capacity. Con-
sequently, both heat capacity, compressibility and capacitive
susceptibility are at a maximum in the melting transition (16,
19). Heat capacity and area compressibility are also related to
the bending elasticity (16, 21, 22), i.e., they are also at a max-
imum in a transition. The bending elasticity, however, is an
important property for the fusion and fission of membranes
(23). This implies that endocytotic and exocytotic events are
potentially enhanced in transitions.
Furthermore, the area compressibility is related to the per-
meability of membranes (24 -- 27). To create pores, the mem-
brane in their surroundings has to be locally compressed,
which is facilitated close to transitions (27, 28). The pores
in the membrane are also related to the formation of lipid ion
channels, i.e., pores in the lipid membranes that display con-
duction patterns that are practically indistinguishable from
protein channels (26, 27, 29 -- 33). The fluctuation-dissipation
theorem implies that large fluctuations go along with longer
fluctuation lifetimes. This results in longer mean open-lifeti-
mes of lipid pores in the transition range (34, 35). The life-
times of lipid membrane-fluctuations span the range from mil-
liseconds to seconds, and are therefore just in the range ob-
served for protein channel open-lifetimes.
It seems likely that also dynamics properties in membra-
nes are related to transitions. For instance, the sound veloc-
ity in lipid dispersions is a function of the membrane com-
pressibility. As a consequence, sound velocities in transitions
are reduced (20, 36). Based on this observation, it has been
proposed that the presence of a phase transition gives rise
to the possibility of the propagation of solitary pulses (soli-
tons) in cylindrical membranes that resemble action poten-
tials (18, 37 -- 39).
Various authors have shown that transitions are influenced
by drugs, e.g., by anesthetics, neurotransmitters or peptides
(35). Anesthetics lower the transition temperature of lipids by
a well-known mechanism called melting-point depression (3,
40 -- 42). The observed pressure-reversal of anesthesia is well
explained by the influence of hydrostatic pressure on melting
transitions (42 -- 44). Within the soliton theory for the nerve
pulse, the effect of anesthetics is explained by the increased
free energy threshold for the induction of a phase transition
(45).
The striking influence of the lipid transition on all kinds of
membrane properties which are of biological relevance sug-
gests a careful evaluation of this phenomenon. In this work
we investigate the melting in various biological membranes,
including the bacteria Escherichia coli and Bacillus subtilis,
lung surfactant, and nerve preparations from rat, sheep and
pig. We show that the transitions in these systems are all
very similar and are found 10-20 degrees below growth or
body temperature. If the growth temperature of bacteria is
changed, the transitions shift as well in the same direction.
We investigate the role of pressure and discuss the lifetimes
of membrane perturbations. In the Discussion, we address the
question of the putative role of such transitions in biology.
Materials and Methods
Lipids: Lipids were purchased from Avanti Polar Lipids
(Birmingham, AL) and used without further purification. Bo-
vine lung surfactant (BLES Biochemicals Inc., London, On-
tario) was a gift from Prof Fred Possmeyer (London, West-
ern Ontario). BLES (bovine lipid extract surfactant) contains
small amounts of membrane soluble proteins (SP-B and SP-
C) and it contains 77 weight% zwitterionic lipids. More than
half of it is dipalmitoyl phosphatidylcholine (DPPC, 41% of
total weight). The exact composition of BLES is given in
(46). CUROSURF (Chiesi Limited, Manchester, UK) was a
gift from Søren Thor Larsen from Haldor Topsoe A/S (Den-
mark). It is prepared from porcine lungsurfactant, and con-
tains around 70-76 weight % zwitterionic phospholipids,
about two-thirds of which is DPPC. The hydrophobic pro-
teins SP-B and SP-C represent about 1% of the total weight.
Details of the composition are given in (47).
Bacterial cells: The E. coli strain XL1 blue with tetracy-
cline resistance (Stratagene, La Jolla, CA) and Bacillus sub-
tilis were grown in an LB-medium at 37 ◦C. Bacterial mem-
branes were disrupted in a French Press at 1200 bar (Gaulin,
APV Homogeniser GmbH, Lubeck, Germany) and centrifu-
ged at low speed in a desk centrifuge to remove solid impuri-
ties. The remaining supernatant was centrifuged at high speed
in a Beckmann ultracentrifuge (50000 rpm) in a Ti70 rotor,
or in a fast desktop centrifuge, to separate the membranes
from soluble proteins and nucleic acids. The pellet was re-
suspended in buffer (33 vol% glycerol plus 67 vol% 10mM
Tris, 1mM EDTA, pH7.2) and centrifuged again. The mem-
brane fractions in the pellets were measured in a calorime-
ter. The concentration of membranes in the calorimetric E.
coli was 26.3 mg/ml as determined from the dry weight of
the samples. Lipid melting peaks and protein unfolding pro-
2
files can easily be distinguished in pressure calorimetry due
to their characteristic pressure dependences, the pressure de-
pendence of lipid transitions being much higher than that of
proteins (17).
Rat central brain: Rat brains were donations from the
Rigshospitalet in Copenhagen (Prof. Niels V. Olsen). They
were kept in a freezer until use. We used the central brain
and parts of the spinal cord. The central brain tissue was
ground with mortar and pestle. The resulting liquid sample
was diluted in a buffer (150 mM KCl, 3mM Hepes, 3 mM
EDTA, pH 7.2-7.4). Subsequently, the sample was sonicated
with a high-power ultrasonic cell disruptor (Branson Ultra-
sonic cell disruptor B15, Danbury, CT, USA) in pulse mode
to prevent heating of the sample. Finally, water soluble parts
were washed away from the sample by centrifugation. The
pellet was assumed to mainly consist of membranes as was
apparent in pressure calorimetry (see results). It was used for
the calorimetric experiments. More details can be found in
(48)
Mammalian spines: Sheep spines were bought from a
local butcher and were kept on ice. The spine was opened
with a saw and the spinal cord was removed. Using scissors
and tweezers the dura mater which surrounds the spinal cord
was removed. The spinal cord was cut into small pieces and
homogenized over a period of 20 minutes with a stator rotor
(Tissue Master 125 W Lab Homogenizer, Omni International,
Inc, Kennesaw, GA) at 33.000 RPM with a 7mm probe head
in 30-second intervals with breaks of 30 seconds to prevent
heating. The homogenate was dissolved in a buffer (150 mM
NaCl, 1mM Hepes, 2mM EDTA, pH 7.4) and spun down in
a desk centrifuge at 3360 g for 15 minutes. The sample was
centrifuged using an MSE Super Minor centrifuge (England)
at 3355 RCF in 15 min intervals. After each round the su-
pernatant was discarded, tubes were filled up to the previous
level with buffer, vortexed and put in the centrifuge for an-
other cycle until the supernatant was completely clear. The
majority of the fibrous tissue that sediments at the bottom was
removed by pouring the viscous pellet into a new tube after
each centrifugation round.
Porcine spines were bought from the local butcher, and a
homogenate of spinal cord tissue was prepared as for sheep
spines. The homogenate was filtrated through a stainless steel
100 mesh with 140 m opening size (Ted Pella, Inc, Redding,
CA) in order to remove fibrous tissue. The homogenate was
dissolved in a 150mM NaCl 11.8 mM phosphate buffer, pH
7.4 and treated as in the sheep spine preparation. Details on
sheep and porcine spine preparations can be found in (49, 50).
Calorimetry: Heat capacity profiles were obtained using
a VP scanning-calorimeter (MicroCal, Northampton, MA) at
scan rates of 20 deg/hr (for spinal cord of rats, sheep, and
chicken) or 30 deg/hr for lung surfactant, E.coli and Bacillus
subtilis membranes. This is much faster than the scan rate
we typically use for pure lipids (typically (cid:54) 5 deg/hr). This is
justified if the expected melting profiles are very broad, and
the cp maximum values are small. A faster scan rate increases
the power of the calorimetric response and thus the strength
of the signal. The small magnitude of the heat capacity leads
to fast relaxation behavior (34, 35) which enables us to scan
fast without hysteresis problems. Pressure calorimetry was
performed in a steel capillary inserted into the calorimetric
cell as previously described (17, 34). In these experiments,
absolute heat capacities are not given. In the porcine spine
preparations, 30 vol% glycerol was added to the sample so-
lution in order to prevent freezing of the sample at tempera-
tures below 0◦C in the calorimeter. A crucial procedure in the
analysis of heat capacity profiles with broad and weak signal
is the subtraction of a baseline. It is shown in the supplemen-
tary information.
Theoretical considerations
In the following sections we outline why a maximum in heat
capacity in a biomembrane melting transition is important for
its physical properties. We show how it determines the mem-
brane compressibility, its elasticity and the lifetime of mem-
brane perturbation which are relevant for the spontaneous
opening of membrane pores.
Through the fluctuation-dissipation theorem, the fluctua-
tions are also coupled to the relaxation time τ,
T 2
L
cp
τ =
(6)
where L ≈ 7 · 108 J·K /mol·s (34, 35). This implies that re-
gions of high heat capacity display slow relaxation processes.
Relaxation times are identical to fluctuation lifetimes (52).
Therefore, it has been suggested that these lifetimes corre-
spond to the open-lifetimes of lipid ion channels, which are
pores in the membrane with a conductance signature that is
indistinguishable from that of a protein ion channel (27, 33).
It has in fact been demonstrated experimentally that the chan-
nel lifetimes are related to the heat capacity of the membrane
for lipid membranes, but also for protein channels reconsti-
tuted into synthetic membranes that display a transition close
to experimental temperature (33, 53).
Fluctuations, susceptibilities and fluctuation life-
times
According to the fluctuation-dissipation theorem, the heat ca-
pacity of a membrane is related to enthalpy fluctuations (16)
through
(cid:18) ∂H
(cid:19)
∂T
cp =
(cid:10)H 2(cid:11) − (cid:104)H(cid:105)2
kT 2
Similarly, the isothermal volume compressibility is related to
volume fluctuations (16)
=
p
(cid:18) ∂V
(cid:19)
T = − 1
κV
(cid:104)V (cid:105)
=
∂p
T
(cid:10)V 2(cid:11) − (cid:104)V (cid:105)2
(cid:104)V (cid:105) · kT
(1)
(2)
Pressure dependence of transitions and volume
compressibility
The pressure dependence of lipid and protein transitions is
very different. Membranes generally increase their transi-
tion temperatures upon increase of hydrostatic pressure be-
cause the excess volume of the membrane is of the order of
+4% and the melting temperature is roughly given by Tm =
(∆E + p∆V )/∆S, i.e., it is proportional to the hydrostatic
pressure. Protein unfolding transitions, in contrast, usually
display a very small excess volume which is negative (54, 55).
A negative excess volume implies that pressure lowers the
unfolding temperature of proteins. Therefore, they can be
denatured by high pressure. This implies that lipid and pro-
tein transitions can be distinguished in calorimetric scans per-
formed at different pressures.
For lipid membrane transitions it has been shown that if
eq. (4) is valid for all temperatures, one can deduce the same
information from heat capacity profiles obtained in the pres-
ence of a hydrostatic pressure difference, ∆p. The enthalpy
(cid:104)∆H(cid:105)∆p
obtained at excess pressure ∆p can be superim-
posed with an enthalpy profile obtained at ∆p = 0 when the
temperature T is rescaled to a new temperature T ∗ according
to the following relations (17):
T
(cid:104)∆H(cid:105)∆p
T
= (1 + γV · ∆p)(cid:104)∆H(cid:105)∆p=0
with T ∗ =
T ∗
≡ f · T ,
T
1 + γV · ∆p
(7)
This relation allows us to check two properties of lipid melt-
ing: 1. a proportional relation between excess volume and
enthalpy holding for all temperatures and 2. the value of γV .
If eq. (7) leads to two superimposable cp profiles, one can
conclude that relations (5) are also valid for biological prepa-
rations. We will use this relation below to determine γV for
a series of biological preparations, and to confirm that the
proportional relation between excess compressibility and heat
capacity also holds for biological melting transitions.
A similar relation for the dependence of membranes on
lateral pressure is very likely but more difficult to measure.
There exists indirect evidence that the proportional relation
3
and the isothermal area compressibility is related to area fluc-
tuations.
(cid:18) ∂A
(cid:19)
=
∂Π
T
(cid:10)A2(cid:11) − (cid:104)A(cid:105)2
(cid:104)A(cid:105) · kT
T = − 1
κA
(cid:104)A(cid:105)
.
(3)
It is a phenomenological finding that for DPPC and some
other lipids (17, 35)
A(T ) ≈ γAH(T )
V (T ) ≈ γV H(T )
;
(4)
where γV = 7.8 · 10−10m2/N (17) and γA = 0.893m/N (16)
for DPPC. It has been shown that the parameters γV and γA
are very similar for different lipids, lipid mixtures and even
biological preparations such as lung surfactant. Using eqs.
(1)-(4), one can conclude that
κV
T =
V · T
γ2
(cid:104)V (cid:105) cp
;
κA
T =
A · T
γ2
(cid:104)A(cid:105) cp
(5)
i.e., the excess compressibilities are proportional to the ex-
cess heat capacity changes. Molecular dynamics simulations
suggest that these relations are also true for absolute heat ca-
pacities and the total compressibilities (51).
Figure 1: Heat capacity profiles of lung surfactant. Left: DPPC large unilamellar vesicles
(LUV). The half width of the peak is 1.7◦. Center: Curosurf, a lipid extract from porcine
lungs. Half width is 14.4◦. Right: Bovine lipid extract surfactant (BLES) as a function of
pressure (top panel) and rescaled using eq.7 (bottom panel). The half width is 15.0 ◦. Both
lung surfactant preparations show a transition peak with a transition temperature of 26.9 ◦C.
between enthalpy and area, and the second relation in eq. (5)
are also correct (16, 21, 22, 51).
Reversibility of protein unfolding and lipid melting
Most proteins unfold irreversibly upon heating, which is a
consequence of the aggregation of unfolded chains that ex-
pose hydrophobic residues to water. Since aggregation is a
slow process, protein unfolding may be partially reversible
on a short time scale or in consecutive scans.
In contrast,
lipid melting is always fully reversible. This allows us to dis-
tinguish protein unfolding from lipid melting in several con-
secutive heating scans in the calorimeter.
Experimental Results
Here, we present studies on various types of cell membranes.
These include two different lung surfactant preparations, E.
coli and B. subtilis membranes, and three different brain and
spine preparations from rat, sheep and pig.
Lung surfactant
Lung surfactant is a lipid film that exists in a monolayer-
bilayer equilibrium on the surface of the alveoli of the lung
(56, 57). Its purpose is to reduce the surface tension at the
air-water interface and to prevent the lung from collapsing.
It contains about 5% of surfactant-associated proteins (SP-A,
B, C and D). The rest is lipids, predominantly DPPC (≈40%)
with a melting temperature of 41.2◦C. In clinical applications,
one often uses lipid extracts from the surfactant. It reduces
the surface tension of the air-water interface in the alveoli
and prevents the lung from collapsing due to the capillary ef-
fect. Two commercial surfactants are bovine lipid extract sur-
factant (BLES) prepared from bovine lungs and Curosurf ex-
tracted from porcine lungs containing about 2% hydrophobic
proteins (46). It consists mostly of phospholipids with minor
contents of the hydrophobic proteins SP-B and SP-C. Since
these preparations are close to native membrane preparations,
4
nearly free of proteins, and available in larger quantities, they
are a good starting material for the study of transitions. Due
to the high content in DPPC, they can readily be compared
to a pure DPPC dispersion. Fig. 1 shows the heat capacity
profiles of DPPC large unilamellar vesicles (LUV) (Fig. 1,
left), Curosurf lung surfactant extract in units of J/g·K (Fig.
1, center) and BLES (Fig. 1,right, top panel) in arbitrary units
at three different pressures (1 bar, 100 bar and 196 bar). The
integral of the cp-profile of Curosurf yields ∆H ≈ 32 J/g
of surfactant. The true value is somewhat higher because
the cp-profile extends to a temperature below zero which is
not accessible in our experiment. For comparison, the major
lipid component of lung surfactant (DPPC) possesses a melt-
ing enthalpy for the main transition of ∆H =45 J/g, which is
of similar order as the surfactant melting enthalpy. Therefore,
we will in the following assume that the transition enthalpies
of DPPC and of lung surfactant are similar.
The right hand panel of Fig. 1 shows the pressure depen-
dence of the transition profile of BLES. The transition maxi-
mum and the half width of the profile at 1 bar are nearly iden-
tical to that of Curosurf. The transition peaks shift towards
higher temperatures upon increasing pressure while maintain-
ing the shape of the profile. One can multiply the absolute
temperature axis with a factor f in order to make the pro-
files recorded at the three different pressures overlap (Fig. 1
right, bottom). The respective factors are f = 0.9925 for the
100 bar recording, and f = 0.985 for the 196 bar recording.
Using eq. (7), one can now determine a value for γV ,
1 − f
f · ∆p
.
γV =
(8)
This calculation yields γV = 7.6 · 10−10 m2/N for the 100
bar measurement and γV = 7.8 · 10−10 m2/N for the 196
bar measurement. This is within error identical to the value
determined for DPPC (γV = 7.8 · 10−10 m2/N) obtained by
Ebel et al. (17) (see this reference also for an estimation of
the errors). This will allow us to make estimates for volume
and area compressibilities of lung surfactant (see below), and
the relaxation times following eqs. (5) and (6).
For Curosurf, we obtained a melting enthalpy compara-
Figure 2: Elastic constants and relaxation time scales
calculated for DPPC large unilamellar vesicles (LUVs,
left) and Curosurf lung surfactant (right).
ble to that of DPPC LUV, and the factor γV was found to be
nearly identical for DPPC LUV and BLES. Taking into ac-
count that additionally the transition maximum and the half
width of the two lung surfactant preparations are nearly iden-
tical, we can make the following assumptions: The melting
enthalpy and the factor γV = 7.8·10−10 m2/J are the same for
the two lung surfactant preparations and DPPC. We assume
that this is also true for the relation between area changes and
enthalpy changes. We therefore estimate that γA = 0.893
m/N for the three preparations, and that the phenomenologi-
cal constant in eq. (6) is given by L = 7 · 108J· K/mol·s. We
further assume that for specific volume and area, the values
for DPPC listed in (16) are reasonably close to the values of
the biological preparation.
We can now calculate the excess volume compressibility,
T , and the (excess) relaxation
T , the area compressibility, κA
κV
times τ in the transition of Curosurf (where the absolute heat
capacity values are known); they are given in Fig. 2 in com-
parison to the DPPC LUV preparation. As shown above, the
compressibilities and the relaxation times are roughly propor-
tional to the excess heat capacity. For the maximum volume
T = 33 · 10−10
compressibility of DPPC LUV we find ∆κV
m2/N, for the area compressibility ∆κT = 22 m/N and for
the relaxation time 0.94 s, respectively. For Curosurf, we
T = 3.6 · 10−10 m2/N, for the area compressibil-
find ∆κV
ity ∆κT = 2.4 m/N and for the relaxation time 0.093 s. This
implies that there is roughly a factor of 10 between DPPC
LUV and lung surfactant, which is also reflected by the find-
ing that the half width of the cp profile is about 10-fold larger
for Curosurf than it is for DPPC LUV.
At physiological temperature, the lipid system (Curosurf)
is above its melting temperature and the excess heat capacity
is close to zero. This implies that the relaxation times in the
lung surfactant preparation are in the millisecond regime. In
the Discussion section we will argue that this time scale is
related to the time scale of lipid ion channels and the time
scale of the nerve impulse.
Escherichia coli and Bacillus subtilis membranes
In a second step, we prepared E. coli and Bacillus subtilis
membranes as indicated in the Methods section. These E.coli
5
cells have a tetracycline resistance, and the cultivation medium
contains tetracycline to prevent growth of other species. The
pelleted membranes were dissolved in a buffer (see Materials
section), filled into a calorimeter and scanned with 30 deg/hr.
In a first experiment, E.coli cells were grown at 37◦C. Fig.
3 (left) shows the first and the second calorimetric scan of the
membranes in the range from 0 to 90 ◦C. The peak attributed
to lipid melting is represented with gray shades. One rec-
ognizes four peaks above the growth temperature (at 48.7◦,
54.5◦, ∼66◦ and ∼88◦C), and one peak below growth tem-
perature (∼23◦C). The peaks above 37◦C become smaller in
consecutive scans. This suggests that they can be attributed to
partially irreversible protein unfolding. The peak at 21.3◦C is
unchanged in the second scan. Its reversible nature suggests
that it can be attributed to the lipid melting transition. The
transition half width is 14.1◦ C, which is nearly the same as
for BLES and Curosurf.
Fig. 3 (right) shows membranes from B. subtilis prepared
in the same manner as the E.coli membranes. The protein
peaks (at 49.7◦, 60.4◦ and 69.8◦C) disappear in the second
upscan due to irreversible unfolding. The lipid transition tem-
perature is about 13.6◦C and the half widths is 17.1 degrees.
This peak is still present in the second complete upscan and
is reversible. Thus, transitions in B. subtilis membranes are
similar to those of the previous biological preparations in re-
spect to half width.
Fig. 3 (center) shows the lipid peak of E.coli membranes
at two different pressures (1 bar and 180 bars). The posi-
tion of the low temperature peak is shifted by 3.45 degrees
towards higher temperature upon application of excess hy-
drostatic pressure. Using eq. (8), one can rescale the tem-
perature axis of the profile measured under pressure such that
it is superimposed on the peak recorded at 1 bar. One finds
that γV = 6.5 · 10−10m2/N, a value that is relatively close
to that obtained for DPPC membranes and lung surfactant
(γV = 7.8 · 10−10m2/N) (17). In addition to the reversibility,
this strongly suggests that the low temperature peak corre-
sponds to lipid melting. As noted above, the excess volume of
most protein unfolding reactions is negative and inconsistent
with the observed pressure dependence of the low tempera-
ture peak. Overall, we find that the melting behavior of E.coli
membranes is very similar to that of lung surfactant both in
Figure 3: Left: Heat capacity scans of E.coli membranes. In the second scan, the protein
peaks are largely reduced. The lipid melting peak (grey shaded) is nearly identical on the first
and the second scan. Center: The lipid melting peak at two different hydrostatic pressures (1
bar and 180 bar, top panel). The bottom panel shows that the cp-profile rescaled according
to eq. (7) yields two nearly superimposable peaks with γV = 6.5 · 10−10n2/N, which is very
similar to the value for artificial lipids (DPPC: 7.8 · 10−10n2/N). Right: First and second
calorimetric scan of B. subtilis membranes. Due to irreversible denaturation, the protein
peaks disappear in the second scan.
respect to width, transition temperature and to the ratio be-
tween excess enthalpy and excess volume. While not show-
ing this here explicitly, we assume that also the total melt-
ing enthalpy per gram of the E.coli lipids is similar to that of
DPPC and lung surfactant, and that one can draw similar con-
clusions with respect to the temperature dependence of the
elastic constants and the time scale of the fluctuations.
In a
further experiment, E.coli cells were grown at three different
temperatures (15◦, 37◦ and 50◦, respectively, see Fig. 4). One
can see that the melting peak of the lipid membrane (shown
in grey shades) moves with the growth temperature. Tg. The
protein unfolding peaks are found at the same positions for
Tg = 15◦C and Tg = 37◦C. This indicates that the cells
adapt to different growth temperatures by changing the lipid
composition. The protein unfolding peaks for Tg = 50◦C are
somewhat different from those at the other two temperatures.
This indicates that we may have selected at high growth tem-
perature a temperature-resistant mutant. Evolution of E. coli
upon exposure to high growth temperatures has in fact been
reported previously (58, 59).
Fig. 5 shows the dependence of the lipid melting transi-
tion on pH and NaCl concentration. Conditions are given in
the figure. Upon lowering the pH from 9 to 5, the transition
temperature increases by 7.6 degrees (Fig. 5, left), which in-
dicates that the E.coli membrane contains a significant frac-
tion of negatively charged lipids. Fig. 5 (right) shows the
lipid melting peak at three different ionic strength conditions.
Increase of the ionic strength from 25 mM NaCl to 300 mM
NaCl leads to a slight decrease in transition temperature.
Figure 4: Left: Adaptation of E.coli membranes to the
growth temperature: 50◦C (a), 37◦C (b) and 15◦C (c).
The lipid melting peak is shifted towards lower temper-
ature upon decreasing the growth temperature. The pro-
tein unfolding peaks in (b) and (c) are identical while they
are slightly different in (a). This indicates that the situa-
tions in (b) and (c) correspond to adaptation rather than
to mutations. Right: Lipid melting peak maximum as a
function of growth temperature.
6
Figure 5: Dependence of melting temperature of E.coli
membranes on pH (left) and ionic strength (right).
Figure 6: Left, top: Heat capacity profiles of porcine spine membranes. The first scan shows
the lipid melting peak and two major protein unfolding peaks. The latter peaks disappear in
the third scan due to irreversible denaturation of the membrane proteins. Left, bottom: Heat
capacity scan of the extracted membrane lipids of the same membrane preparation. Center:
Melting profile of rat central brain membranes. The lipid membrane peak is conserved in
the second scan while the protein unfolding is irreversible and the respective protein peaks
disappear. From (48). Right: Heat capacity profiles of sheep spine membranes. From (49,
50).
Nerve membranes
In the electromechanical theory for the action potential (18,
37, 39, 60), the melting transition plays a central role. There-
fore, the investigation of melting profiles of nerve membranes
is of particular importance. We have chosen central brain
and spinal cord tissues because we assume that here most
membranes are related to signal conduction. However, we
cannot distinguish different membranes. Our results repre-
sent an average over all types of membranes in the tissue.
Membranes display a larger density than do soluble proteins.
Therefore, our procedure attempts to separate soluble parts
from the membranes. The membrane fraction analyzed here
contains all membranes from the cells including those of or-
ganelles. We assume that a major fraction of the membranes
are myelin sheets and axonal membranes.
Fig. 6 shows the heat capacity profiles for three differ-
ent membranes preparations from the spinal cord of pig, the
central brain (medulla and parts of the spinal cord) from rat
and the spinal cord of sheep. The left panel shows the heat
capacity profiles of porcine spine membranes. It displays a
maximum at 23.1◦C. There are two further peaks associated
with protein unfolding located above physiological tempera-
ture that disappear in the second and third complete upscan
due to irreversible denaturation. The lipid peak in the third
scan seems to be somewhat affected by the denaturation of
the proteins, displaying a maximum at about 29◦C. Proteins
aggregating upon denaturation thus diminish lipid-protein in-
teractions. This would fit to the finding that extracted lipids
display a cp maximum at roughly the same temperature (bot-
tom trace of the left panel). In this heat capacity profile one
also finds a very sharp low enthalpy peak at around 51.5◦C.
Such a sharp peak is typical for vesicles composed of a pure
lipid component. Its origin is not clear and we disregard it
due to its low enthalpy content. The center panel of Fig. 6
shows the membrane preparation from rat central brain. Here,
we find a lipid membrane melting peak at around 28.2◦C,
and two major protein unfolding peaks at 59◦C, with a mi-
nor shoulder at 50.2◦C, and 79.9◦C that disappear in the sec-
ond complete heating scan due to irreversible denaturation of
proteins. The right hand panel of Fig. 6 shows the calorimet-
ric profile of sheep spine that displays a lipid melting maxi-
mum at 24.1◦C) and protein unfolding peaks at 49.4, 56.7 and
83.3◦C). The protein peaks also disappear in a consecutive
heating scan (data not shown because we could not identify
a reasonable baseline). We also investigated small amounts
of chicken spine where similar calorimetric events as shown
here for the other nerve preparations could be identified (data
not shown). Summarizing one can state that the protein peaks
are at similar positions in the three preparations, whereas the
lipid peaks are not at exactly the same position. This may
be an artifact due to heterogeneity of the biological sample
preparations.
The half widths of the lipid peaks are of similar order as in
the E. coli, B. subtilis and lung surfactant preparations. Even
though we have not been able to measure absolute heat capac-
ities due to the lack of knowledge of the amount of lipids in
the sample, we assume that the overall magnitude of the heat
capacity and the enthalpy changes are of comparable order as
in lung surfactant (see discussion).
lipid
protein
protein
protein
pig
23.1
-
56.8
78.8
rat
28.2
50.2
59.0
79.7
sheep
24.1
49.4
56.7
82.3
Table 1: Major peaks in the heat capacity profiles of na-
tive nerve preparations in the first complete heating scan.
7
Discussion
We showed here that in all biological membrane preparations
we have investigated in this study, one finds lipid melting
transitions that occur about 10-15 deg below physiological
or growth temperature. The preparations were from biologi-
cal sources as different as Gram-negative and Gram-positive
bacteria, lung surfactant and nerve membranes. In a recent
publication, we have shown that such transitions also occur
in cancer cells of various origins (61).
The lipid transitions can be identified by various indica-
tors:
• In repeated scans over a large temperature interval the
protein unfolding peaks are mostly irreversible while
lipid melting transitions are reversible.
• Lipid transitions display a very different excess pres-
sure dependence as compared to protein unfolding.
While the excess volume of lipid transitions is positive,
leading to an increase of the transition temperature of
the order of 1 deg/40 bars (Figs. 1 and 3), the excess
volume of proteins is usually negative indicating that
the temperature of denaturation for proteins is lowered
upon increasing pressure (54, 55).
• Comparison to melting profiles of lipid extracts (see
Fig. 6) shows that the melting peaks in the native prepa-
rations are similar with respect to transition tempera-
ture and transition half width.
The enthalpy of the protein unfolding transitions is of similar
order as that of lipid melting, although exact ratios depend on
the quality of the membrane preparations and the degree to
which soluble proteins can be washed out of the samples.
We investigated various membrane preparations: two pre-
parations of lung surfactant, membranes from the Gram-
negative bacterium E.coli, characterized by an inner cytoplas-
mic membrane and an outer cell membrane, and the Gram-
positive bacterium B. subtilis, displaying a cytoplasmic lipid
membrane and an outer peptidoglycan layer, and three prepa-
rations from central brain and spinal cord nerves (pig, sheep
and rat). In all of these preparations, we found lipid melt-
ing transitions in a range between 10-20 degrees below stan-
dard growth or body temperatures. They all share some com-
mon features, such as displaying a transition width of about
10-15◦ and calorimetric events being detectable up to phys-
iological temperature. Thus, the physiological temperature
in all of the preparations was found to lie just between the
lipid transition and the protein unfolding transitions such that
minor perturbations of the membranes will move the mem-
branes into the transition regime. It seems likely that the tran-
sition slightly below body temperature is a generic feature of
most cell membranes, and this phenomenon may serve im-
portant purposes in the functioning of a cell. This feature is
conserved from unicellular organisms like bacteria up to the
complex nerve system of mammals.
For two of the biological preparations and the pure lipid
DPPC, we measured the pressure dependence of the lipid
transition. The pressure dependence yields a coefficient relat-
ing the excess volume and the excess enthalpy of the transi-
tions, ∆V (T ) = γV ∆H(T ) . We have shown previously that
8
T = (γ2
the temperature dependence of the excess volume is propor-
tional to that of the excess enthalpy, ∆V (T ) = γV ∆H(T )
(17). The constant γV was found to be 7.8 · 10−10 m2/N for
DPPC (16, 17), 7.6 − 7.8 · 10−10 m2/N for the lung surfac-
tant preparation (BLES) and 6.5 · 10−10 m2/N for E. coli.
A value close to this was confirmed by Pedersen et al in MD-
simulations even outside of the transition for DPPC (62). This
implies that the correlation between volume and enthalpy also
holds outside of the transition regime. While these values
are subject to some experimental deviations for the biolog-
ical preparations as reflected by the broad transition peaks
and uncertainties of baseline subtraction, they are reasonably
close for DPPC, lung surfactant and E. coli membranes. Fur-
ther, we found that the overall excess enthalpy of the transi-
tion of lung surfactant was similar to that of DPPC. These
two facts have an important consequence:
the excess vol-
ume compressibility can now be determined from the heat
capacity profile when the overall enthalpy of the transition
and the magnitude of the excess heat capacity are known:
V T /V )·∆cp (16). This implies that the membrane
∆kV
volume is more compressible in the melting transition. We
assume the same proportionality to be true for the membrane
area (excluding the membrane part of proteins): ∆A(T ) =
γA · ∆H(T ), with γA = 0.893 m2/J for DPPC (16). Thus,
we can calculate the changes in area compressibility from the
AT /A) · ∆cp (16). The latter
heat capacity to be ∆kA
correlation also allows determination of changes in the bend-
ing elasticity (the inverse bending modulus in Helfrich's the-
(cid:113)
ory) to be κB = (16/D2)·∆cp (16). The heat capacity is also
correlated to the sound velocity. The velocity of sound in the
S is
membrane plane is given by c =
the adiabatic area compressibility. It is related to the isother-
T − (T /V cp) · (dV /dTp)2.
mal compressibility via κA
Here, both κA
T , the volume expansion coefficient are simple
functions of the heat capacity. Therefore, the sound velocity
can be determined from the heat capacity profile. In previous
work, we demonstrated that the volume compressibility and
sound velocity in lipid dispersion can be calculated from the
heat capacity (16, 20, 63). By analogy, we assume this to be
correct in the membrane plane, where the sound velocity in
the fluid membrane is given by 176 m/s (18). It is signifi-
cantly slower in the transition regime.
1/ρA · κA
S , where κA
T = (γ2
S = κA
Our findings have important consequences as summarized
below. Many of them have been confirmed for synthetic lipid
membranes.
• at the transition temperature, the volume and the lateral
compressibility are at maximum, i.e., the membranes
are more compressible within the transition (16, 64)
• the bending elasticity is at maximum, i.e., membranes
in the transition range are much more flexible than in
the fluid or gel membrane (16, 22)
• membrane lifetimes such as open dwell-time of mem-
brane pores or curvature fluctuation lifetimes are slower
in the transition
• the sound velocity is in a minimum (20, 36)
• the magnitude of the effects is similar for all biological
preparations shown here because they display similar
shapes of their transitions, both with respect to transi-
tion temperature and half width of the transition.
• since thickness and area of membranes change, transi-
tions can also affect the capacitance of the membrane
(19, 65). We have previously shown that the capaci-
tance in an artificial membrane can double when going
from the gel to the fluid phase (66).
There is good reason to assume that the above effects also ex-
ist for biological membranes, but they will be less pronounced
because the heat capacity at maximum is lower and the width
of the transition is considerably larger. For the membrane
function this implies that
• the probability for membrane fusion and fission events
is enhanced because it depends on curvature elasticity
(23)
• membrane pores are more abundant because their en-
ergy depends on the lateral compressibility (27, 28)
• lateral sound pulses called solitons can be generated in
membranes (18)
• anesthetics, hydrostatic pressure or pH shift the tran-
sition and thereby generate changes in compressibility,
bending elasticity, pore formation probability (26, 27,
32), open lifetime of membrane pores (32) and soli-
ton excitability (45). For instance, it has been shown
that anesthetics can reduce the open probability of lipid
channels in artificial membranes in the complete ab-
sence of proteins (26, 29).
Our data on E.coli membranes show that the lipid melt-
ing peak adapts to the growth temperature.
If the latter is
higher, the lipid transition also moves upwards in tempera-
ture in order to maintain a certain distance from physiolog-
ical temperature. Adaptation of the transition temperature
in E. coli membranes has already been reported by (13). It
seems likely that the adaptation consists of a change in the
fraction of lipids with high and low melting temperatures, or
the ratio of saturated to unsaturated chains (67). In partic-
ular, for trout livers the lipid composition was shown to be
different in winter and summer (68). At 20◦C, the fraction
of saturated lipids is higher, whereas at 5◦C the fraction of
poly-unsaturated lipids is higher. The composition of lipid
membranes also adapts to the presence of solvents (among
those: acetone, chloroform or benzene) (69) or to hydrostatic
pressure in deep sea bacteria (70). High pressure increases
the melting transition temperature. Consequently, the frac-
tion of unsaturated lipids increases because these lipids dis-
play lower melting temperatures which opposes the pressure-
effect. For this reason it seems from the few experimental
studies that the relative position of the melting transition rela-
tive to physiological temperature is a property actively main-
tained by the organisms.
During a melting transition, membranes display a coexis-
tence of gel and fluid phases that are easy to observe in flu-
orescence microscopy. This has been well documented for
artificial membranes (4, 5, 71, 72), monolayers (73 -- 76) and
also for lung surfactant (77, 78). Phase coexistence is obvious
from fluorescence microscopy at the transition temperature,
and domains are large. Domain coexistence has frequently
been discussed in connection with so-called lipid rafts (79).
Rafts are rich in sphingomyelin, cholesterol and certain pro-
teins (80 -- 83). Sphingolipids are known to display high melt-
ing temperatures, which are further enhanced by cholesterol.
It is therefore tempting to assume that lipid rafts are gel do-
mains swimming in a fluid environment. The fact that phys-
iological temperature is at the upper end of the lipid melting
transitions suggests that domains at physiological conditions
must be small. This may be the origin of the difficulties to
identify rafts at physiological conditions.
If rafts are indi-
cators for lipid melting processes, is likely that rafts will be
larger and much easier to detect 10-20 degrees below body
temperatures.
We and others have shown in the past that the melting
transition of synthetic membranes shifts towards lower tem-
peratures (3, 40, 42, 84) upon application of general and local
anesthetics. This phenomenon follows the simple freezing-
point depression law that relates the concentration of anes-
thetic in the fluid phase to the shift in transition tempera-
ture. This simple law has two advantages: 1. It is consis-
tent with the famous Meyer-Overton correlation (42, 85) and
2. it is drug-unspecific, meaning that it does not depend on
the particular chemistry of the anesthetic molecules. In con-
nection with the present investigation we found that the anes-
thetic pentobarbital lowers transition temperatures in ovine
and porcine spine by 2 to 3 degrees when exposed to a buffer
with up to 20 mol% pentobarbital (data not shown). A shift of
about 3 degrees was found for 8 mM pentobarbital on DPPC
membranes (3). Thus, the effect reported here is of similar
order of magnitude. Since our biological preparations are
subject to preparational irreproducibilities and the additional
effect of sample aging, we take this as a strong indication that
the effect of anesthetics in biological preparations is similar
to that in artificial membranes. We will investigate this ef-
fect more carefully in future studies. Alltogether, our results
are consistent with the finding that alcohols from ethanol to
decanol lower the critical temperature of plasma membranes
from rat leukemia cells (86).
An important consequence of the heat capacity maximum
is the accompanying increase in the elastic susceptibilities,
leading to more flexible and compressible membranes but
also to decrease in the sound velocity along membrane cylin-
ders. This temperature and density dependence of the lat-
eral sound velocity is the key element of the soliton theory
for nerves that reproduces many properties of action poten-
tials (18, 37 -- 39, 60). In this model for signal propagation,
the nerve pulse is a solitary pulse in which the lateral density
and membrane thickness increases transiently. Both effects
have been found in experiments (87 -- 90). Recently, a theory
was put forward that explains how anesthetics can change the
stimulation threshold for these density pulses (45).
Conclusions
We have shown in this work that the lipid melting transitions
exist in various biological membranes in vitro. They influ-
ence various mechanical features, which as a consequence
9
influence membrane properties such as vesicle fusion prob-
ability, pore formation probabilities and their lifetimes, and
the generation of solitary pulses in axonal membranes. Since
the position of a transition depends on the presence of drugs,
pH, pressure, or the presence of proteins, nature gains a pow-
erful tool to influence cell surface properties via the control
of macroscopic thermodynamic properties of the biological
membrane as a whole.
Author contributions: TM and FT prepared sheep and pig
membranes from spinal cord and investigated them calori-
metrically, SM prepared and investigated rat membranes from
central brain, DP performed some of the calorimetric E.coli
and B.subtilis measurements, MK prepared all E.coli and
B.subtilis membranes. TH performed E.coli, B.subtilis and
lung surfactant experiments and wrote the article. The article
was proofread and commented by the other authors.
Acknowledgments: M. Konrad acknowledges continuous
support by the Max-Planck-Institute for Biophysical Chem-
istry. We thank Niels Olsen from Rigshospitalet Copenhagen
for donations of rat brains. We thank Søren Thor Larsen (Hal-
dor Topsoe A/S, Technical University of Denmark and Uni-
versity of Copenhagen) for a donation of Curosurf lung sur-
factant, and Fred Possmeyer for a donation of lung surfactant
preparations (BLES). A few data sets were used in a different
context before. Two of the scans in Fig. 4 were used in (1).
The scans in the right panel of Fig. 1 were used in (17). We
added them for completeness of the discussion. This work
was supported by the Villum foundation (VKR 022130).
References
1. Heimburg, T., 2007. Thermal biophysics of membranes. Wiley
VCH, Berlin, Germany.
2. Lee, A. G. 1977. Lipid phase transitions and phase diagrams.
II. Mixtures involving lipids. Biochim. Biophys. Acta 472:285 --
344.
3. Graesbøll, K., H. Sasse-Middelhoff, and T. Heimburg. 2014.
The thermodynamics of general and local anesthesia. Biophys.
J. 106:2143 -- 2156.
4. Korlach, J., J., P. Schwille, W. W. Webb, and G. W. Feigenson.
1999. Characterization of lipid bilayer phases by confocal mi-
croscopy and fluorescence correlation spectroscopy. Proc. Natl.
Acad. Sci. USA 96:8461 -- 8466.
5. Bagatolli, L. A., and E. Gratton. 1999. Two-photon fluores-
cence microscopy observation of shape changes at the phase
transition in phospholipid giant unilamellar vesicles. Biophys.
J. 77:2090 -- 2101.
6. Steim, J. M., M. E. Tourtellotte, J. C. Reinert, R. N. McElhaney,
and R. L. Rader. 1969. Calorimetric evidence for the liquid-
crystalline state of lipids in a biomembrane. Proc. Natl. Acad.
Sci. USA 1963:104 -- 109.
7. Reinert, J. C., and J. M. Steim. 1970. Calorimetric detection
of a membrane-lipid phase transition in living cells. Science
168:1580 -- 1582.
8. Ashe, G. B., and J. M. Steim. 1971. Membrane transitions in
gram-positive bacteria. Biochim. Biophys. Acta 233:810 -- 814.
9. Wisnieski, B. J., J. G. Parkes, Y. O. Huang, and C. F. Fox. 1974.
Physical and physiological evidence for two phase transitions in
cytoplasmic membranes of animal cells. Proc. Natl. Acad. Sci.
USA 71:4381 -- 4385.
10. Chow, E. I., S. Y. Chuang, and P. K. Tseng. 1981. Detection of
a phase-transition in red-cell membranes using positronium as
a probe. Biochim. Biophys. Acta 646:356 -- 359.
11. Haest, C. W., A. J. Verkleij, J. de Gier, R. Scheek, P. Verver-
gaert, and L. L. M. van Deenen. 1974. The effect of lipid
phase transitions on the architecture of bacterial membranes.
Biochim. Biophys. Acta 356:17 -- 26.
12. Nag, K., J. Perez-Gil, M. L. F. Ruano, L. A. D. Worthman,
J. Stewart, C. Casals, and K. K. M. W. Keough. 1998. Phase
transitions in films of lung surfactant at the air-water interface.
Biophys. J. 74:2983 -- 2995.
13. Nakayama, H., T. Mitsui, M. Nishihara, and M. Kito. 1980.
Relation between growth temperature of E. coli and phase tran-
sition temperatures of its cytoplasmic and outer membranes.
Biochim. Biophys. Acta 601:1 -- 10.
14. White, S. H., and T. E. Thompson. 1973. Capacitance, area,
and thickness variations in thin lipid films. Biochim. Biophys.
Acta 323:7 -- 22.
15. Jamieson, G. A., and D. M. Robinson, 1977. Mammalian cell
membranes, volume 2. Butterworth, London.
16. Heimburg, T. 1998. Mechanical aspects of membrane ther-
modynamics. Estimation of the mechanical properties of lipid
membranes close to the chain melting transition from calorime-
try. Biochim. Biophys. Acta 1415:147 -- 162.
17. Ebel, H., P. Grabitz, and T. Heimburg. 2001. Enthalpy and vol-
ume changes in lipid membranes. i. the proportionality of heat
and volume changes in the lipid melting transition and its im-
plication for the elastic constants. J. Phys. Chem. B 105:7353 --
7360.
18. Heimburg, T., and A. D. Jackson. 2005. On soliton propaga-
tion in biomembranes and nerves. Proc. Natl. Acad. Sci. USA
102:9790 -- 9795.
19. Heimburg, T. 2012. The capacitance and electromechanical
coupling of lipid membranes close to transitions. the effect of
electrostriction. Biophys. J. 103:918 -- 929.
20. Schrader, W., H. Ebel, P. Grabitz, E. Hanke, T. Heimburg,
M. Hoeckel, M. Kahle, F. Wente, and U. Kaatze. 2002. Com-
pressibility of lipid mixtures studied by calorimetry and ultra-
sonic velocity measurements.
J. Phys. Chem. B 106:6581 --
6586.
21. Evans, E., and R. Kwok. 1982. Mechanical calorimetry of large
dimyristoylphosphatidylcholine vesicles in the phase transition
region. Biochemistry 210.
22. Dimova, R., B. Pouligny, and C. Dietrich. 2000. Pretransitional
effects in dimyristoylphosphatidylcholine vesicle membranes:
Optical dynamometry study. Biophys. J. 79:340 -- 356.
23. Kozlovsky, Y., and M. M. Kozlov. 2002. Stalk model of mem-
brane fusion: Solution of energy crisis. Biophys. J. 82:882 -- 895.
24. Papahadjopoulos, D., K. Jacobson, S. Nir, and T. Isac. 1973.
Phase transitions in phospholipid vesicles. fluorescence polar-
ization and permeability measurements concerning the effect of
temperature and cholesterol. Biochim. Biophys. Acta 311:330 --
340.
25. Sabra, M. C., K. Jørgensen, and O. G. Mouritsen. 1996. Lin-
dane suppresses the lipid-bilayer permeability in the main tran-
sition region. Biochim. Biophys. Acta 1282:85 -- 92.
26. Blicher, A., K. Wodzinska, M. Fidorra, M. Winterhalter, and
T. Heimburg. 2009. The temperature dependence of lipid mem-
brane permeability, its quantized nature, and the influence of
anesthetics. Biophys. J. 96:4581 -- 4591.
27. Heimburg, T. 2010. Lipid ion channels. Biophys. Chem. 150:2 --
22.
10
28. Nagle, J. F., and H. L. Scott. 1978. Lateral compressibility
of lipid mono- and bilayers. Theory of membrane permeability.
Biochim. Biophys. Acta 513:236 -- 243.
29. Wodzinska, K., A. Blicher, and T. Heimburg. 2009. The ther-
modynamics of lipid ion channel formation in the absence and
presence of anesthetics. blm experiments and simulations. Soft
Matter 5:3319 -- 3330.
30. Wunderlich, B., C. Leirer, A. Idzko, U. F. Keyser, V. Myles,
T. Heimburg, and M. Schneider. 2009. Phase state depen-
dent current fluctuations in pure lipid membranes. Biophys. J.
96:4592 -- 4597.
31. Laub, K. R., K. Witschas, A. Blicher, S. B. Madsen,
A. Luckhoff, and T. Heimburg. 2012. Comparing ion conduc-
tance recordings of synthetic lipid bilayers with cell membranes
containing trp channels. Biochim. Biophys. Acta 1818:1 -- 12.
32. Blicher, A., and T. Heimburg. 2013. Voltage-gated lipid ion
channels. PloS ONE 8:e65707.
33. Mosgaard, L. D., and T. Heimburg. 2013. Lipid ion channels
and the role of proteins. Acc. Chem. Res. 46:2966 -- 2976.
34. Grabitz, P., V. P. Ivanova, and T. Heimburg. 2002. Relaxation
kinetics of lipid membranes and its relation to the heat capacity.
Biophys. J. 82:299 -- 309.
35. Seeger, H. M., M. L. Gudmundsson, and T. Heimburg. 2007.
How anesthetics, neurotransmitters, and antibiotics influence
the relaxation processes in lipid membranes. J. Phys. Chem.
B 111:13858 -- 13866.
36. Halstenberg, S., T. Heimburg, T. Hianik, U. Kaatze, and R. Kri-
vanek. 1998. Cholesterol-induced variations in the volume and
enthalpy fluctuations of lipid bilayers. Biophys. J. 75:264 -- 271.
37. Heimburg, T., and A. D. Jackson. 2007. On the action poten-
tial as a propagating density pulse and the role of anesthetics.
Biophys. Rev. Lett. 2:57 -- 78.
38. Andersen, S. S. L., A. D. Jackson, and T. Heimburg. 2009.
Towards a thermodynamic theory of nerve pulse propagation.
Progr. Neurobiol. 88:104 -- 113.
39. Lautrup, B., R. Appali, A. D. Jackson, and T. Heimburg. 2011.
The stability of solitons in biomembranes and nerves. Eur. Phys.
J. E 34:57.
40. Kaminoh, Y., S. Nishimura, H. Kamaya, and I. Ueda. 1992. Al-
cohol interaction with high entropy states of macromolecules:
critical temperature hypothesis for anesthesia cutoff. Biochim.
Biophys. Acta 1106:335 -- 343.
41. Kharakoz, D. P. 2001. Phase-transition-driven synaptic exocy-
tosis: A hypothesis and its physiological and evolutionary im-
plications. Biosci. Rep. 210:801 -- 830.
42. Heimburg, T., and A. D. Jackson. 2007. The thermodynamics
of general anesthesia. Biophys. J. 92:3159 -- 3165.
43. Trudell, J. R., D. G. Payan, J. H. Chin, and E. N. Cohen.
1975. The antagonistic effect of an inhalation anesthetic and
high pressure on the phase diagram pf mixed dipalmitoyl-
dimyristoylphosphatidylcholine bilayers. Proc. Natl. Acad. Sci.
USA 72:210 -- 213.
44. Kamaya, H., I. Ueda, P. S. Moore, and H. Eyring. 1979. Antago-
nism between high pressure and anesthetics in the thermal phase
transition of dipalmitoyl phosphatidylcholine bilayer. Biochim.
Biophys. Acta 550:131 -- 137.
45. Wang, T., T. Muzi´c, A. D. Jackson, and T. Heimburg. 2018.
The free energy of biomembrane and nerve excitation and
the role of anesthetics. Biochim. Biophys. Acta 1860:doi:
10.1016/j.bbamem.2018.04.003.
46. Zhang, H., Q. Fan, Y. E. Wang, C. R. Neal, and Y. Y.
Zuo. 2011. Comparative study of clinical pulmonary surfac-
tants using atomic force microscopy. Biochim. Biophys. Acta
1808:1832 -- 1842.
47. Taeusch, H. W., K. Lu, and D. Ramierez-Schrempp. 2002. Im-
proving pulmonary surfactants. Acta Pharm. Sin. 23:11 -- 15.
48. Madsen, S. B., 2011. Thermodynamics of nerves. Master's
thesis, University of Copenhagen.
49. Tounsi, F., 2015. The correlation between critical ananesthetic
dose and melting temperatures in synthetic membranes. With
comments on the possible implications of the soliton model on
epilepsy. Master's thesis, University of Copenhagen.
50. Muzic, T., 2016. The effect of anesthetics on phase transitions
in biological membranes. Master's thesis, Niels Bohr Institute,
University of Copenhagen.
51. Pedersen, U. R., G. H. Peters, T. B. Schrøder, and J. C. Dyre.
2010. Correlated volume-energy fluctuations of phospholipid
membranes: A simulation study. J. Phys. Chem. B 114:2124 --
2130.
52. Onsager, L. 1931. Reciprocal relations in irreversible processes.
II. Phys. Rev. 38:2265 -- 2279.
53. Seeger, H. M., A. Alessandrini, and P. Facci. 2010. KcsA re-
distribution upon lipid domain formation in supported lipid bi-
layers and its functional implications. Biophys. J. 98:371a.
54. Royer, C. A. 2002. Revisiting volume changes in pressure-
induced protein unfolding. Biochim. Biophys. Acta 1595:201 --
209.
55. Ravindra, R., and R. Winter. 2003. On the temperature - pres-
sure free-energy landscape of proteins. Chem. Phys. Chem.
4:359 -- 365.
56. Perez-Gil, J., and T. E. Weaver. 2010. Pulmonary surfactant
pathophysiology: Current models and open questions. Physiol-
ogy 25:132 -- 141.
57. Echaide, M., C. Autilio, R. Arroyo, and J. Perez-Gil. 2017.
Restoring pulmonary surfactant membranes and films at the res-
piratory surface. Biochim. Biophys. Acta 1859:1725 -- 1739.
58. Rudolph, B., K. M. Gebendorfer, J. Buchner, and J. Winter.
2010. Evolution of Escherichia coli for growth at high tem-
peratures. J. Biol. Chem. 285:19029 -- 19034.
59. Guyot, S., L. Pottier, A. Hartmann, M. Ragon, J. H. Tiburski,
P. Molin, E. Ferret, and P. Gervais. 2013. Extremely rapid
acclimation of Escherichia coli to high temperature over a few
generations of a fed-batch culture during slow warming. Micro-
biologyOpen 3:52 -- 63.
60. Heimburg, T., and A. D. Jackson, 2008. Thermodynamics of the
nervous impulse. In Structure and Dynamics of Membranous
Interfaces. (Nag, K., editor). Wiley, 317 -- 339.
61. Højholt, K. L., T. Muzi´c, S. D. Jensen, M. Bilgin, J. Nyland-
sted, T. Heimburg, S. K. Frandsen, and J. Gehl. 2019. Cal-
cium electroporation and electrochemotherapy for cancer treat-
ment: Importance of cell membrane composition investigated
by lipidomics, calorimetry and in vitro efficacy. Sci. Rep.
9:4758.
62. Pedersen, R.
J., 2011.
Electrophysiological measure-
ments of spontaneous action potentials in crayfish nerve
in relation to the soliton model. Master's thesis, Uni-
versity
http://membranes.nbi.dk/thesis-
pdf/2011 Masters RolfPedersen.pdf.
of Copenhagen.,
63. Mosgaard, L. D., A. D. Jackson, and T. Heimburg.
2013.
Fluctuations of systems in finite heat reservoirs with applica-
tions to phase transitions in lipid membranes. J. Chem. Phys.
139:125101.
64. Evans, E. A. 1974. Bending resistance and chemically induced
moments in membrane bilayers. Biophys. J. 14:923 -- 931.
65. Mosgaard, L. D., K. A. Zecchi, and T. Heimburg.
2015.
Mechano-capacitive properties of polarized membranes. Soft
Matter 11:7899 -- 7910.
66. Zecchi, K. A., L. D. Mosgaard, and T. Heimburg.
2017.
Mechano-capacitive properties of polarized membranes and the
application to conductance measurements of lipid membrane
patches. J. Phys.: Conf. Ser. 780:012001.
11
89. Iwasa, K., I. Tasaki, and R. C. Gibbons. 1980. Swelling of nerve
fibres associated with action potentials. Science 210:338 -- 339.
90. Gonzalez-Perez, A., L. D. Mosgaard, R. Budvytyte, E. Villa-
gran Vargas, A. D. Jackson, and T. Heimburg. 2016. Solitary
electromechanical pulses in lobster neurons. Biophys. Chem.
216:51 -- 59.
67. Hazel, J. R. 1995. Thermal adaptation in biological membranes:
Is homeoviscous adaptation the explanation? Ann. Rev. Phys-
iol. 57:19 -- 42.
68. Hazel, J. R. 1979. Influence of thermal acclimation on mem-
brane lipid composition of rainbow trout liver. Am. J. Physiol.
Regulatory Integrative Comp. Physiol. 287:R633 -- R641.
69. Ingram, L. O. 1977. Changes in lipid composition of Es-
cherichia coli resulting from growth with organic solvents and
with food additives. Appl. Environ. Microbiol. 33:1233 -- 1236.
70. DeLong, E. F., and A. A. Yayanos. 1985. Adaptation of the
membrane lipids of a deep-sea bacterium to changes in hydro-
static pressure. Science 228:1101 -- 1103.
71. Baumgart, T., S. T. Hess, and W. W. Webb. 2003. Imaging coex-
isting fluid domains in biomembrane models coupling curvature
and line tension. Nature 425:821 -- 824.
72. Hac, A., H. Seeger, M. Fidorra, and T. Heimburg. 2005. Diffu-
sion in two-component lipid membranes -- a fluorescence corre-
lation spectroscopy and monte carlo simulation study. Biophys.
J. 88:317 -- 333.
73. Losche, M., E. Sackmann, and H. M. hwald. 1983. A flu-
orescence microscopic study concerning the phase diagram of
phospholipids. Ber. Bunsenges. Phys. Chem. 87:848 -- 852.
74. McConnell, H. M., and V. T. Moy. 1988. Shapes of finite two-
dimensional lipid domains. J. Phys. Chem. 92:4520 -- 4525.
75. Knobler, C. M. 1990. Seeing phenomena in flatland: studies of
monolayers by fluorescence microscopy. Science 249:870 -- 874.
76. Gudmand, M., M. Fidorra, T. Bjørnholm, and T. Heimburg.
2009. Diffusion and partitioning of fluorescent lipid probes in
phospholipid monolayers. Biophys. J. 96:4598 -- 4609.
77. de la Serna, J. B., J. Perez-Gil, A. C. Simonsen, and L. A.
Bagatolli. 2004. Cholesterol rules: direct observation of the
coexistence of two fluid phases in native pulmonary surfac-
tant membranes at physiological temperatures. J. Biol. Chem.
279:40715 -- 40722.
78. de la Serna, J. B., G. Oradd, L. Bagatolli, A. C. Simonsen,
D. Marsh, G. Lindblom, and J. Perez-Gil. 2009. Segregated
phases in pulmonary surfactant membranes do not show coex-
istence of lipid populations with differentiated dynamic proper-
ties. Biophys. J. 97:1381 -- 1389.
79. Almeida, P. F. 2014. The many faces of lipid rafts. Biophys. J.
106:1841 -- 1843.
80. Brown, D. A., and E. London. 1998. Functions of lipid rafts in
biological membranes. Ann. Rev. Cell Develop. Biol. 14:111 --
136.
81. Bagnat, M., S. Keranen, A. Shevchenko, and K. Simons. 2000.
Lipid rafts function in biosynthetic delivery of proteins to the
cell surface in yeast. Proc. Natl. Acad. Sci. USA 97:3254 -- 3259.
82. Bagnat, M., and K. Simons. 2002. Lipid rafts in protein sort-
ing and cell polarity in budding yeast saccharomyces cerevisiae.
Biol. Chem. 383:1475 -- 1480.
83. Edidin, M. 2003. The state of lipid rafts: From model mem-
branes to cells. Ann. Rev. Biophys. Biomol. Struct. 32:257 -- 283.
84. Johnson, S., and K. W. Miller. 1970. Antagonism of pressure
and anaesthesia. Nature 228:75 -- 76.
85. Overton, C. E., 1991. Studies of narcosis. Chapman and Hall,
New York. English version of 'Studien der Narkose' from 1901.
86. 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 -- -2759.
87. Tasaki, I., K. Kusano, and M. Byrne. 1989. Rapid mechanical
and thermal changes in the garfish olfactory nerve associated
with a propagated impulse. Biophys. J. 55:1033 -- 1040.
88. Iwasa, K., and I. Tasaki. 1980. Mechanical changes in squid
giant-axons associated with production of action potentials.
Biochem. Biophys. Research Comm. 95:1328 -- 1331.
12
Supplement
Baseline removal
Figure S1: Baseline correction procedure. Left: Scan of E.coli membranes without baseline
correction, and subjectively chosen baseline for the first and the second scan, and the baseline
corrected scans (bottom). Center: Baseline subtraction for porcine spine membranes. Right:
Baseline subtraction for sheep spinal cord membranes.
The determination of a calorimetric baseline is difficult for weak signals. It is to a certain degree subjective because calorimetric events
span from the lowest temperature to the highest temperature (typically from 0 degrees to 95 degrees centigrade) and there are no obvious
reference point in the traces as it is the case for single lipid traces. Especially at the lowest and highest temperature the cp-values are
probably subject to high error. It cannot generally be expected that the peaks always begin at the lowest calorimetric temperature and they
may extend into the negative temperature regime. It is therefore helpful to use the lowest possible temperature for the start of an up-scan,
as done below for the porcine spine membranes, which were recorded in the presence of glycerol. We have tried to be as conservative as
possible with the removal of a baseline. Examples are given in Fig. S1. For the given reasons one should consider the calorimetric traces
rather as a qualitative rather than as exact measures.
13
|
1210.7480 | 1 | 1210 | 2012-10-28T17:32:52 | Noisy NFkB oscillations stabilize and sensitize cytokine signaling in space | [
"physics.bio-ph",
"q-bio.CB"
] | NF-kB is a major transcription factor mediating inflammatory response. In response to pro-inflammatory stimulus, it exhibits characteristic response -- a pulse followed by noisy oscillations in concentrations of considerably smaller amplitude. NF-kB is an important mediator of cellular communication, as it is both activated by and upregulates production of cytokines, signals used by white blood cells to find the source of inflammation. While the oscillatory dynamics of NF-$\kappa$B has been extensively investigated both experimentally and theoretically, the role of the noise and the lower secondary amplitude has not been addressed.
We use a cellular automaton model to address these issues in the context of spatially distributed communicating cells. We find that noisy secondary oscillations stabilize concentric wave patterns, thus improving signal quality. Furthermore, both lower secondary amplitude as well as noise in the oscillation period might be working against chronic inflammation, the state of self-sustained and stimulus-independent excitations.
Our findings suggest that the characteristic irregular secondary oscillations of lower amplitude are not accidental. On the contrary, they might have evolved to increase robustness of the inflammatory response and the system's ability to return to pre-stimulated state. | physics.bio-ph | physics | Noisy NF-κB oscillations stabilize and sensitize cytokine signaling in space.
Sirin W. Gangstad1,2, Cilie W. Feldager1,2, Jeppe Juul1, and Ala Trusina1
1Niels Bohr Institute, University of Copenhagen and
2These authors contributed equally to this work.
NF-κB is a major transcription factor mediating inflammatory response.
In response to pro-
inflammatory stimulus, it exhibits characteristic response -- a pulse followed by noisy oscillations
in concentrations of considerably smaller amplitude. NF-κB is an important mediator of cellular
communication, as it is both activated by and upregulates production of cytokines, signals used by
white blood cells to find the source of inflammation. While the oscillatory dynamics of NF-κB has
been extensively investigated both experimentally and theoretically, the role of the noise and the
lower secondary amplitude has not been addressed.
We use a cellular automaton model to address these issues in the context of spatially distributed
communicating cells. We find that noisy secondary oscillations stabilize concentric wave patterns,
thus improving signal quality. Furthermore, both lower secondary amplitude as well as noise in the
oscillation period might be working against chronic inflammation, the state of self-sustained and
stimulus-independent excitations.
Our findings suggest that the characteristic irregular secondary oscillations of lower amplitude are
not accidental. On the contrary, they might have evolved to increase robustness of the inflammatory
response and the system's ability to return to pre-stimulated state.
INTRODUCTION
The regulatory network of NF-κB is an important
constituent of the immune system as it regulates hun-
dreds of genes in response to extracellular stimuli such as
pathogens, cytokines and stress [1, 2]. These responses
include apoptosis, cell proliferation, and inflammatory
response [1, 3].
During inflammatory response, the NF-κB network is
activated by an increase in extracellular concentration
of cytokines; small signaling molecules commonly used
in intercellular communication. Once activated, NF-κB
up-regulates the cells' own cytokine production, thereby
amplifying the external signal and passing it to the neigh-
bouring cells [4, 5] either by diffusion or through gap-
junctions -- channels formed by physically interacting
cells [5].
When passing the signal from one cell to the next in
this manner, the tissue acts as an excitable medium. Re-
cent theoretical research has shown that propagating ele-
vated cytokine concentrations as waves through this "ex-
citable tissue" might represent an optimal way of passing
the signal to the blood vessels, where the cytokines are
absorbed [6, 7]. When neutrophils (white blood cells) de-
tect the cytokines in the blood stream, they will follow
the cytokine concentration gradient back to the site of
infection [8], in order to destroy or contain the cause of
the infection [9].
Single cell measurements [4, 10] and modeling ap-
proaches [11, 12] revealed that the concentration of active
NF-κB oscillates with an initial high amplitude peak and
several consecutive lower amplitude peaks. The experi-
ments also revealed a considerable amount of noise in
this response, and simulations suggest that the noise is
in fact induced by an inherent component of the network
[13 -- 16]. The effect of noise on the wave propagation of
cytokines through the tissue has not yet been elucidated,
but the presence of a noise-inducing component suggests
that the noise itself may play a key role in the immune
response. This has motivated us to ask the following
questions: How does noise affect wave propagation? Is
the system equally sensitive to irregularities in the pe-
riod of oscillations as to irregularities in the refractory
period? Does noise contribute to the onset of chronic
inflammation -- the state where waves are self-sustained
and do not depend on stimulus?
THE MODEL
The tissue is modeled as an excitable media using
a parallel cellular automaton, an algorithm frequently
used as a mathematical idealization of biological self-
organizing systems [17 -- 21]. Our model comprises 101 ×
101 cells placed on a square grid. All cells are initially
inactive, but once activated by their neighbouring cells
they start producing cytokines in an irregularly oscillat-
ing fashion with an initial high peak followed by several
lower peaks (see Fig. 1a). The amplitude of the initial
peak is normalized to 1 and the mean oscillation period
is estimated from current literature to be 100 minutes
[13], corresponding to 12 time steps in the model. The
amplitude A of all secondary peaks and the relative size
ηo of the gaussian noise in the oscillation period are kept
as free parameters, but retained similar to values exper-
imentally measured in nuclear NF-κB.
A cell is activated if the combined cytokine production
of its 8 nearest neighbours, the Moore-neighbourhood,
exceeds a threshold θ = 3. That is, a cell can be ac-
tivated if 3 of its neighbours are in their first peak of
cytokine production, or if e.g. 2 neighbours are in their
2
1
0
2
t
c
O
8
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
0
8
4
7
.
0
1
2
1
:
v
i
X
r
a
2
RESULTS
An example of the effect of noise in the wave propa-
gation is shown in Fig. 1. Here, noise is added to the
period of oscillations (Fig. 1F-I) and to the duration of
the refractory period (Fig. 1J-M). The addition of noise
has two biologically interesting effects on the system:
Higher secondary amplitude renders the system
more robust to noise.
In order to destroy or contain pathogens, the neu-
trophils need a distinct concentration gradient to follow
to the site of infection. Adding sufficient amount of noise
to the period of oscillations and/or in the length of the re-
fractory period of each cell causes the propagating waves
to disintegrate or to lose their radial gradient. Biolog-
ically, this may compromise the effectiveness of the in-
flammatory response.
The time it takes the system to lose its wave struc-
ture depends on the amount of noise in the system. In
order to examine this time before breakdown, the state
of the system over time is compared to the state of the
system the first time the waves fill the entire grid. This
is achieved by calculating the Pearson correlation coeffi-
cient between cytokine concentration of the first grid in
which the waves have reached the outer edges and the
grid in all later time steps:
C1,t = (cid:104) g1 − (cid:104)g1(cid:105)
σ1
· gt − (cid:104)gt(cid:105)
σt
(cid:105).
(1)
Here, g1 is the cytokine levels of the first grid that is
filled with waves, and gt is the grid at a later time step.
Angle brackets denote spacial average and σ denotes spa-
tial standard deviation. The correlation coefficients are
plotted as a function of time in Fig.2A-B. The correla-
tion coefficients decay due to noise destroying the waves.
The characteristic time for breakdown of wave structure
is defined as the last time the correlation coefficient to
the first wave is more than 0.25 (see Fig. 2A-B). The
exact value of this threshold is not essential for the main
results, as long as it is low enough to allow for some ir-
regularity in wave patterns.
The correlation method allows us to investigate the
characteristic time for breakdown as a function of noise
in the period the oscillations and noise in the refractory
period. Interestingly, the secondary amplitude A, has a
significant effect on the systems ability to sustain wave
propagation. As is seen in Fig. 2C-E, the parameter
range with stable wave patterns increases with increas-
ing secondary amplitude. Thus, high secondary ampli-
tude makes the system more robust to noise. One pos-
sible explanation for this is that secondary waves might
FIG. 1. A: The oscillatory cytokine production of a single
cell subject to constant stimulation. The amplitude of the
initial peak is normalized to 1 while all secondary peaks have
amplitude A. When the cell leaves refractory period, it is
here immediately re-excited to its initial peak. B-E: show
wave propagation in an ideal system without any noise. F-I:
have ηo = 40 % noise in the oscillation period and no noise in
the refractory period, ηr = 0. J-M: show the system with no
noise in the oscillation period, ηo = 0, but ηr = 10 % noise in
the refractory period. The secondary amplitude is A = 0.25
in all panels A-M. In the time series F-I and J-M, snapshots
are taken at t = 20, 60, 100 and 200 hours. We see that both
kinds of noise gradually generate spirals of active cells, which
may mislead neutrophils and cause chronic inflammation.
first peak and 1/A other neighbours are in a secondary
peak. When a cell is activated, it enters a refractory
period in which it is completely unresponsive to the lo-
cal cytokine concentration, and will produce cytokines at
its own pace. The mean refractory period is estimated
from current literature to be 200 minutes [22], but here
is also added a gaussian noise of relative size ηr, kept
as a free parameter. Once out of the refractory period,
the cell becomes susceptible to input signals again, and
if the local cytokine concentration should rise above the
threshold, the cell will re-initiate the cytokine produc-
tion, responding with the initial high peak followed by
several low peaks (see Fig. 1a).
A stimulus of 3×3 constantly excited cells in the middle
of the grid was added to act as a pathogen and initiate a
train of cytokine waves propagating through the tissue.
Refractory PeriodOscillationPeriodTime1ACytokine Production0.00.51.0Cytokine ConcentrationBCDEFGHIJKLMAReactivation3
Noise leads to self-sustained excitations resembling
chronic inflammation.
To further address the issue of chronic inflammation,
we tried removing the stimulus after 120 periods of os-
cillations, corresponding to 200 hours. If this eliminated
the cytokine production, such that no cell in the tissue
was re-excited to its first peak after another 48 periods of
oscillations, we defined the system to be sensitive to the
removal of the stimulus. If not, the system was declared
insensitive. A system that is insensitive to the removal of
the stimulus is not able to relax back to the pre-infected
state, and is reminiscent of chronic inflammation.
We found that noise in the oscillation period and/or in
the duration of refractory period can create stable addi-
tional sources and spirals, which excite the system inde-
pendently of the stimulus, see Fig. 3A-D. Thus, noise not
only destroys the radial gradient of the wave pattern, so
that neutrophils will be unable to locate the stimulus, but
the system can become insensitive so that removing the
stimulus will not terminate the inflammatory response.
As the system is particularly sensitive to noise in the
refractory period, we investigate how the insensitivity de-
pends on the level of refractory noise, ηr. We define the
probability of becoming insensitive to the stimulus, P (I),
as the fraction out of 30 runs, where the system becomes
insensitive.
Interestingly, with no noise in the oscillation pe-
riod, the probability to become insensitive is a non-
monotonous function of refractory noise, as seen in see
Fig. 3E. With no noise, the system is sensitive, P (I) = 0
at ηr = 0, but a small increase in noise has a dramatic
effect and completely destroys the sensitivity, P (I) ≈ 1
when ηr = 0.01. At this small noise level, the insensitiv-
ity is caused by a double spiral very close to the center,
where the stimulus once was, as seen in Fig. 3A-B. To
sustain this double spiral, the previous wavefront repeat-
edly excites the central cells through a narrow passage
of cells, as illustrated in Fig. 3B. If the noise level is in-
creased further to 0.04 < ηr < 0.1, this signal can not be
transmitted through the narrow passage, and the prob-
ability that the system is insensitive decreases. This ex-
plains the decline in probability for insensitivity in figure
3E. Increasing noise further, ηr > 0.1 produces spirals
in rich numbers everywhere in the tissue, as shown in
Fig. 3C-D. These spiral structures are quite stable and
can continue to excite cells on long timescales after the
stimulus has been removed.
To better understand the non-monotonous behavior
of the systems sensitivity, we have tested the case with
larger system size, which should give higher chance for
spirals to originate away from the center. As expected,
Fig. 3F shows that the onset of permanent insensitiv-
ity has been expedited to lower noise levels, reducing the
range of stimulus sensitivity. Furthermore, adding noise
to the oscillation period prevents the formation of double
FIG. 2. A-B: Correlation to the first wave as a function
of time in hours, illustrating how the characteristic time for
breakdown of wave propagation is found. The red line cor-
responds to the correlation threshold of 0.25. A shows cor-
relations of a system with little noise that is able to sustain
waves. Parameters: A = 0.25, ηo = 0.02% and ηr = 0.01%.
B shows correlations of a system with sufficient noise to break
the waves. Parameters: A = 0.25, ηo = 0% and ηr = 0.05%.
C-E: Characteristic time for breakdown as function of noise
levels in the period of oscillations and in the length of refrac-
tory period for different secondary amplitudes. The secondary
amplitude in panel C-E is 0.15, 0.25, and 0.35, respectively.
help straighten up the lagging sections of the major wave-
fronts by increasing the cytokine concentration at the in-
flections points and effectively accelerating them. Since
lagging cells can get excited by having e.g. two neigh-
bours in primary peak and 1/A neighbours in secondary
peak, a high secondary amplitude will lead to a larger
acceleration. There is, however, an upper limit to the
secondary amplitude: Waves should only originate from
the central stimulus, so the combined secondary ampli-
tudes from the 8 nearest neighbours of a cell should never
exceed the activation threshold θ = 3. That is, if the am-
plitude of secondary oscillations is larger than θ/8, cells
have a possibility of getting activated by the secondary
oscillations of its neighbours. This could happen any-
where in the tissue, not just close to a stimulus, causing
an inflammatory response from healthy tissue leading to
chronic inflammation. Although this upper limit on the
secondary amplitude is unknown for living cells, it is in-
teresting to note that the experimentally observed sec-
ondary amplitude is significantly smaller than the initial
peak, with the secondary amplitude being about 20 % of
the first peak [13]!
05010015000.10.20.30.400.020.040.060.080.1ηr00.10.20.30.400.10.20.30.4ηoηoηoTime for breakdownCDE00.251500501000.500.75-0.25ABCorrelation with first waveTime Time15005010020040060080010001200Breakdown4
Fig. 3G). This also explains why a small noise in the
oscillation period drastically extends the characteristic
time before breakdown of wave propagation in Fig. 2D-
E.
DISCUSSION
The inflammatory response is initiated by the NF-κB
regulatory network. The dynamics of NF-κB response is
very peculiar when measured on single cell level. While
the oscillatory dynamics of NF-κB in single cells have
been the focus of extensive research [4, 13], little atten-
tion has been given to the the role of the lower secondary
amplitude and the noise in oscillations.
Rather than seeking the single-cell explanations, we
have taken the perspective of spatially distributed cells,
which collectively propagate cytokine signals through the
tissue, thereby recruiting white blood cells from the blood
stream to the site of infection. This approach led us to
several interesting findings: a) The system is more sensi-
tive to noise in refractory period than in oscillation pe-
riod. b) Moderate noise in secondary oscillations stabi-
lizes concentric wave patterns in the presence of noise in
the refractory period. The effect is stronger with higher
secondary amplitude. However, the secondary amplitude
should never be so high that a group of cells in their sec-
ondary oscillation can re-excite a cell to its first peak, as
this potentially would create numerous sources of prop-
agating waves in the system. This may help to explain
why the experimentally observed secondary amplitude is
indeed much smaller than the initial peak, ∼ 20%. This
prediction can be tested experimentally. The propaga-
tion of waves could be measured in a 2D culture of mam-
malian cells with a fluorescent reporter fused to NF-kB
[4]. In this case nuclear NF-kB will serve as a proxy for
cytokine production. The secondary amplitude is con-
trolled by A20, such that higher expression of A20 re-
sults in lower amplitude. There already exist cell lines
where the amounts of A20 can be tuned externally [23].
Combining the two: fluorescent microscopy of 2D culture
of cells with externally tunable A20, one can test our
predictions that higher amplitude would stabilize wave
propagation.
c) Increasing noise in refractory period increases the
chance of self-sustained excitations, thus decoupling the
excitations from the original source. This situation is in
effect very similar to chronic inflammation -- inflamma-
tion which persists and re-occurs even after the source of
damage has been removed [24]. The system can show a
very rich, non-monotonous behavior in how it turns in-
sensitive with increasing refractory noise. Remarkably,
noise in secondary oscillations postpones the onset of in-
sensitivity, thus rendering system more robust to low
noise in refractory period. Our findings related to the
effect of noise in the refractory period are limited to the
FIG. 3. A-D show two distinct ways the system can be-
come insensitive to the stimulus. A-B: For small noise levels,
a double spiral appears close to the removed stimulus. To
sustain this double spiral, the previous wavefront needs to
excite the central cells through a narrow passage. Parame-
ters: A = 0.25, ηo = 0%, ηr = 2%. C-D: For larger noise
levels spirals arise numerous places in the tissue. Parameters:
A = 0.25, ηo = 0%, ηr = 20%. E: The probability P (I) that
the noise in the refractory period causes the system to become
insensitive to removal of stimulus as a function of noise in the
refractory period. Interestingly, the curve is non-monotonic.
F: Same as E but with a grid size of 202 × 202 cells instead
of 101 × 101. The larger system size expedites the emergence
of stable spirals. G: Same as E but with with a noise in the
oscillation period of ηo = 10 %, which destabilizes the dou-
ble spirals for low noise levels, making the insensitivity curve
monotonic.
spirals close to the center, as the cytokine signal is again
prevented to pass through the narrow passage of Fig. 3B.
As a result, noise in the period of oscillation of ηo = 10%
turns P (I) into a monotonously increasing function (see
0.00.51.0Cytokine Concentrationηr00.10.20.30.4ABCDE010.5010.5Narrow passageGηr00.10.2Fηr00.10.2PIPIinflammatory response alone, but are general for other
excitable media models.
Overall, our investigation shows that both of the ex-
perimentally observed features: the lower amplitude of
secondary oscillations and the noise in the period of os-
cillations, are crucial for stabilizing wave patterns as well
as maintaining system sensitivity.
Aknowledgement This study was supported by the
Danish National Research Foundation through the Cen-
ter for Models of Life and Steno fellowship (AT).
[1] A. R. Brasier, Cardiovasc Toxicol 6, 111 (2006).
[2] A. Hoffmann and D. Baltimore, Immunol Rev 210, 171
(2006).
[3] H. L. Pahl, Oncogene 18, 6853 (1999).
[4] T. K. Lee et al., Sci Signal 2, ra65 (2009).
[5] C. A. Kasper et al., Immunity 33, 804 (2010).
[6] P. Yde, M. H. Jensen, and A. Trusina, Phys Rev E Stat
Nonlin Soft Matter Phys 84, 051913 (2011).
[7] P. Yde, B. Mengel, M. H. Jensen, S. Krishna, and
A. Trusina, BMC Syst Biol 5, 115 (2011).
[8] J. Geiger, D. Wessels, and D. R. Soll, Cell Motil Cy-
toskeleton 56, 27 (2003).
[9] V. Witko-Sarsat, P. Rieu, B. Descamps-Latscha,
5
P. Lesavre, and L. Halbwachs-Mecarelli, Lab Invest 80,
617 (2000).
[10] S. Tay et al., Nature 466, 267 (2010).
[11] B. Mengel, S. Krishna, M. H. Jensen, and A. Trusina,
Physica A: Statistical Mechanics and its Applications
391, 100 (2012).
[12] T. Lipniacki, P. Paszek, A. R. A. R. Brasier, B. Luxon,
and M. Kimmel, J Theor Biol 228, 195 (2004).
[13] L. Ashall et al., Science 324, 242 (2009).
[14] M. Marcello and M. R. H. White, Biochem Soc Trans
38, 1247 (2010).
[15] D. A. Turner et al., J Cell Sci 123, 2834 (2010).
[16] T. Lipniacki and M. Kimmel, Cardiovasc Toxicol 7, 215
(2007).
[17] S. Wolfram, Rev. Mod. Phys. 55, 601 (1983).
[18] G. Bub, A. Shrier, and L. Glass, Phys Rev Lett 88,
058101 (2002).
[19] G. Bub, A. Shrier, and L. Glass, Phys Rev Lett 94,
028105 (2005).
[20] S. Sawai, P. A. Thomason, and E. C. Cox, Nature 433,
323 (2005).
[21] C. Marr and M.-T. Hutt, Physics Letters A 349, 302
(2006).
[22] D. E. Nelson et al., Science 306, 704 (2004).
[23] S. L. Werner, D. Barken, and A. Hoffmann, Science 309,
1857 (2005).
[24] C. Nathan and A. Ding, Cell 140, 871 (2010).
|
1911.02527 | 1 | 1911 | 2019-11-01T16:21:38 | Body-terrain interaction affects large bump traversal of insects and legged robots | [
"physics.bio-ph",
"cs.RO",
"q-bio.QM"
] | Small animals and robots must often rapidly traverse large bump-like obstacles when moving through complex 3-D terrains, during which, in addition to leg-ground contact, their body inevitably comes into physical contact with the obstacles. However, we know little about the performance limits of large bump traversal and how body-terrain interaction affects traversal. To address these, we challenged the discoid cockroach and an open-loop six-legged robot to dynamically run into a large bump of varying height to discover the maximal traversal performance, and studied how locomotor modes and traversal performance are affected by body-terrain interaction. Remarkably, during rapid running, both the animal and the robot were capable of dynamically traversing a bump much higher than its hip height (up to 4 times the hip height for the animal and 3 times for the robot, respectively) at traversal speeds typical of running, with decreasing traversal probability with increasing bump height. A stability analysis using a novel locomotion energy landscape model explained why traversal was more likely when the animal or robot approached the bump with a low initial body yaw and a high initial body pitch, and why deflection was more likely otherwise. Inspired by these principles, we demonstrated a novel control strategy of active body pitching that increased the robot maximal traversable bump height by 75%. Our study is a major step in establishing the framework of locomotion energy landscapes to understand locomotion in complex 3-D terrains. | physics.bio-ph | physics | Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Body-terrain interaction affects large bump traversal of insects and legged robots
Sean W. Gart and Chen Li
Department of Mechanical Engineering, Johns Hopkins University
3400 N. Charles St, 126 Hackerman Hall, Baltimore, Maryland 21218-2683, USA
E-mail: [email protected]
Keywords: obstacle, cockroach, Blaberus discoidalis, bio-inspiration, locomotion energy landscape,
terradynamics
Abstract
Small animals and robots must often rapidly traverse large bump-like obstacles when moving
through complex 3-D terrains, during which, in addition to leg-ground contact, their body inevitably comes
into physical contact with the obstacles. However, we know little about the performance limits of large
bump traversal and how body-terrain interaction affects traversal. To address these, we challenged the
discoid cockroach and an open-loop six-legged robot to dynamically run into a large bump of varying height
to discover the maximal traversal performance, and studied how locomotor modes and traversal
performance are affected by body-terrain interaction. Remarkably, during rapid running, both the animal
and the robot were capable of dynamically traversing a bump much higher than its hip height (up to 4 times
the hip height for the animal and 3 times for the robot, respectively) at traversal speeds typical of running,
with decreasing traversal probability with increasing bump height. A stability analysis using a novel
locomotion energy landscape model explained why traversal was more likely when the animal or robot
approached the bump with a low initial body yaw and a high initial body pitch, and why deflection was
more likely otherwise. Inspired by these principles, we demonstrated a novel control strategy of active body
pitching that increased the robot's maximal traversable bump height by 75%. Our study is a major step in
1
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
establishing the framework of locomotion energy landscapes to understand locomotion in complex 3-D
terrains.
1. Introduction
The ability to traverse large obstacles is crucial for both animals and robots. Many insects [1 -- 5],
reptiles [6 -- 9], small birds [10], and small mammals [11] encounter large bump-like obstacles in their natural
habitats, such as branch litter and roots on the rainforest floor and rock beds near river or in desert
environments (figure 1), which they need to traverse in order to survive [12, 13]. Similarly, the speed at
which search-and-rescue robots [14] traverse terrains such as building rubble and landslides where large
bumps are abundant (figure 1) could determine the success of failure of a critical mission [15].
Figure 1. Examples of bump-like obstacles in complex 3-D terrains: (a) leaf and branch litter, (b) rock
beds, (c) exposed roots, and (d) building rubble. All images were available under Creative Commons CC0
at pixabay.com.
2
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Previous studies of animal locomotion over bump-like obstacles focused on how animals control,
adjust, or plan body and leg motions using sensory information [1, 2, 6, 16, 17] to traverse the obstacle,
during which body collision with the obstacle is often being actively avoided. Such avoidance of body
collision is possible either because the animal starts from at rest or moves slowly enough (lower than 30%
of walking-to-running transition speed) [1, 2, 18, 19], and/or because the obstacle is small enough (up to
twice the animal's hip height) [8, 20, 21]. Similarly, empirical adjustment of leg control [22] and sensor-
based planning [23] have enabled rapid-running legged robots to traverse bump-like obstacles such as stairs,
small walls, and blocks, typically up to the robot's hip height [24 -- 28]. Although some robots could traverse
a bump-like obstacle beyond one hip height high [24, 29, 30], doing so was realized by ballistic jumping
[22, 26, 30] which requires stopping to plan leg movement.
However, during dynamic locomotion in natural and artificial environments, legged animals and
robots can sometimes rapidly run into bump-like obstacles even higher than their hip height. In this case,
because sensory and neuromuscular delays [31] make it challenging to plan and adjust body and leg
movements substantially and rapidly enough [3, 32], avoidance of body collision and physical interaction
with obstacles may also be impossible [3, 5, 33 -- 35]. It is not well understood how well animals and robots
can traverse bump-like obstacles much larger than their hip height, and whether the often inevitable body-
terrain interaction during dynamic locomotion over these obstacles affects traversal.
Unlike locomotion on simpler 2-D surfaces where an animal can walk, run, or climb for extended
time, when negotiating large obstacles in such complex 3-D terrains, animals more often use and transition
between diverse locomotor modes [6, 19, 35]. Previous studies focused on how animals use sensory
information to make decisions about which locomotor mode to use. For example, cockroaches walking at
less than 10 cm s−1 either climb over, tunnel underneath [19], follow, or turn away from an obstacle [5, 36]
depending on how its antennae contacted the obstacle. Recent studies revealed that, during rapid
locomotion, the passive physical interaction between animal/robot body with the terrain can also play a
major role in initiating diverse locomotor modes while traversing complex 3-D terrains and determining
3
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
traversal probability, even when no sensory information is available [35]. Our companion paper also
demonstrated that passive body-terrain interaction strongly affected traversal performance of large a gap-
like obstacle comparable to body size [37].
Inspired by these recent discoveries, we hypothesized that passive body-terrain interaction, if
appropriate, can also help animals and robots traverse bump-like obstacles much higher than their hip
height. To test this, we observed how well the discoid cockroach and a cockroach-inspired, open-loop
legged robot traversed a large bump up to 4 times the hip height, and studied how locomotor modes and
traversal performance depended on bump height and body-terrain interaction during bump encounter. Based
on these observations, we used a novel modeling framework of locomotion energy landscapes [35] to begin
to understand why body-terrain interaction affected locomotor modes and traversal performance. In
addition, we compared animal and robot results to gain insights into the role of active sensory feedback.
Finally, we demonstrated a control strategy to enhance large bump traversal during dynamic locomotion
used our robot with an active tail [37].
2. Methods
The majority of methods follow those described in the companion paper [37]. Below we summarize
methods used in this study that differed from the companion study.
2.1. Experiments
For animal experiments, we used 15 male Blaberus discoidalis cockroaches (Pinellas County
Reptiles, St Petersburg, FL, USA). The animals measured 4.5 ± 0.8 cm in length and 2.4 ± 0.2 cm in width
and weighed 2.7 ± 0.3 g. We used the obstacle track described in our companion study [37], but replaced
the gap with a bump that spanned the entire width of the track (figure 2(c)). We tested four bump heights,
4
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
H = 0.5, 1, 1.5, and 2 cm, ranging from 1 to 4 times the animal's hip height (zhip = 0.5 cm) and collected a
total of 553 accepted trials.
To quantify the large variety of locomotor modes during bump traversal, we used an obstacle track
capable of high-throughput semi-automated data collection [38]. To automatically stimulate the animal's
escape response and encourage them to run over the bump, we attached 3-D printed pushing pads to linear
actuators (Progressive Automation, Richmond, BC, CAN) that moved up and down repeatedly to motivate
the animal to run from shaded enclaves at each end of the track. The enclaves were shaded from four 500
W work lamps (Coleman Cable, Waukegan, IL, USA) illuminating the middle test section of the track. We
constructed a 4 cm long bump with 3-D printed blocks (Ultimaker 2+, Geldermalsen, Netherlands) and
covered them with white paper cardstock (Pacon 4-ply railroad poster board, Appleton, WI, USA) to
achieve uniform friction properties.
Four synchronized cameras (JAI-5000-PCML, Copenhagen, DEN) filmed the test area from the
dorsal view at 50 frames s−1 at a resolution of 2560 × 2048 pixels and a 500 µs shutter time. To track the
motion of the animal, we attached 10 mm × 10 mm BEEtags [39] dorsally above the body center of mass
(CoM). Because tags were occasionally pulled off by pushing pads, we removed the animal's wings before
attaching the tags. This did not have an impact on their locomotion because the animal did not contact the
bump obstacle with the dorsal surface of its body.
For robot experiments, we tested four bump heights, H = 2.5, 5, 7.5, and 10 cm, ranging from 1 to
4 times the robot's hip height (zhip = 2.5 cm). For each of the four bump heights, we tested four different
running speeds (17 ± 7 cm s−1, 39 ± 11 cm s−1, 61 ± 16 cm s−1, and 70 ± 13 cm s−1) and three initial body
heading angles (0°, 30°, and 60°). We collected 5 trials each for each combination of running speed, initial
body heading angle, and bump height, resulting in a total of 240 trials. We constructed the bumps using 30
cm × 60 cm stacked acrylic sheets. To prevent leg slip when the robot started to move, we covered the
bottom surface leading to the bump with 50 grit sandpaper. To reduce the friction between the head of the
robot and the front face of the bump, we covered the bump surface with polystyrene. For the robot's
5
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
anterior, we chose to use an ellipsoidal shape similar to the animal head shape to allow direct comparison
between animal and robot observations.
For the tailed robot experiment to test our control hypothesis, we tested the tailed robot at a constant
speed of 70 ± 3 cm s−1 with a fixed initial body heading perpendicular to the bump. For both with and
without tail actuation, we varied bump height from 1.5 hip height to 4 hip height with an increment of 0.5
hip height. We collected 10 trials for each hip height, resulting in a total of 120 trials.
Figure 2. Experimental setup and definition of geometric and kinematic variables. (a) Discoid cockroach,
hip height zhip = 0.5 cm, body length b = 4.5 ± 0.8 cm. (b) Legged robot, hip height zhip = 2.5 cm, body
length b = 25 cm. We used BEEtags [39] to measure 3-D position (x, y, z) and orientation (yaw α, pitch β,
and roll γ) of the body. (c) Schematic of the bump obstacle. Bump height H was varied from 1 -- 4 hip height
and bump length l = 0.8 body length for animal experiment and l = 1 body length for the robot experiment.
Two dorsal cameras and one lateral camera were used to record experiments.
2.2. Data analysis
For both animal and robot experiments, we obtained locomotor mode transition ethograms [16, 19,
40] for each traversal attempt. In experiments, we observed four different locomotor modes (see section
6
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
3.2): climbing, deflection, simultaneous climbing and deflection, which we referred to as slanted climbing,
and flipping over (which only occurred for the robot).
To determine the instantaneous locomotor mode, we measured kinematic data (velocity, body pitch
β, and body yaw α) using a moving average over a window of three frames. Transition between modes
occurred when two adjacent frames had different locomotor modes. We obtained initial kinematic
measurements such as approach speed v0, initial body pitch β0, and initial body yaw α0 by averaging data
from two video frames prior to bump contact. We observed only a small difference between the velocity
heading and body yaw immediately prior to gap encounter (7° ± 8° for the animal, 2° ± 8° for the robot).
There we assumed velocity heading always equaled body yaw. Further definitions of these metrics are
described in our companion study [37].
To automatically categorize the animal and robot's locomotion during bump negotiation in one of
the four modes observed, we quantitatively defined them as follows.
(1) Climbing: For all except the 1 hip height bump, the straight ascent mode occurred when body
pitch β exceeded 30° and body yaw was less than 50°. For the 1 hip height bump, because upward body
pitch β rarely exceeded 30°, we assumed that if the body was not deflected, the animal or robot traversed
in the straight ascent mode.
(2) Deflection: We defined deflection to have occurred when body pitch β was smaller than 30°
and either of three criteria were satisfied: body yaw α exceeded 50°; the ratio of lateral velocity to forward
velocity exceeded 0.9 at any time after bump collision; or if initial body yaw α0 was greater than 50° and
body yaw α increased more than 5° at any time during the bump traversal attempt. These three deflection
criteria were necessary because the animal and robot occasionally bounced off of the bump, moving
laterally without turning its body yaw α in the lateral direction. Alternatively, it encountered the bump with
a high initial body yaw α0 that did not increase during traversal. In this case, the body was moving straight
and was not deflected in its heading.
7
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
(3) Slanted climbing: The animal and robot climbed with a high body yaw α and velocity heading
angle. This mode occurred when body pitch β exceeded 30° and body yaw α exceeded 50° simultaneously.
This mode was usually a brief transitional mode between straight ascent and deflection.
(4) Flipping over: Occasionally, the robot body pitched or rolled over 90° and landed on its dorsal
side. We only observed this mode in the robot and the animal never flipped over during traversal of bumps
up to 4 hip height.
We defined a trial to result in traversal if the center of mass (CoM) crossed the leading edge of the bump.
To check if the animal adjusted its kinematics after antennae contact, we measured approach speed,
body pitch, and body yaw when the antennae touched and when the head touched the front face of the
bump. We found no statistically significant changes in approach speed and body pitch (P = 0.113, P =
0.995, respectively, repeated-measures ANOVA) and only a slight increase in body yaw from 21° ± 15° to
25° ± 18° (P = 0.0002, repeated-measures ANOVA).
To analyze locomotor pathways for the animal experiment, we first calculated the transition
probability between each pair of locomotor modes for all trials from one individual. Then to compare
between bump sizes, we averaged the means of all individuals for each bump size. For the robot, we
calculated the transition probably between locomotor modes by averaging all trials for each bump height.
To further highlight the differences in body orientation between these two locomotor pathways, we
averaged head height, body pitch, and body yaw for trials that showed only the straight ascent or only the
deflection mode. We excluded diagonal ascent because it was a transitional mode.
To measure bump traversal performance, we measured traversal speed, defined as the average
forward speed of the body from when the head reached the near edge of the bump to when the center of
mass reached the near edge of the bump.
8
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
3. Results and Discussion
3.1. Animal and robot are capable of traversing a large bump dynamically
Running at an average approach speed of 41 ± 15 cm s−1 (independent of bump height, P = 0.132,
repeated-measures ANOVA), the animal was capable of traversing all the bumps tested, up to the largest 4
hip height bump (figure 3(a), filled circles). Remarkably, the animal was capable of dynamically traversing
a large bump at speeds higher than its walking-to-running transition speed with decreasing probability with
bump height (figure 3(a), open circles) [41, 42] (traversal speed > 33 cm s−1, Froude number Fr > 1.5).
Additionally, the animal could traverse a bump twice as high as found in previous studies, with a traversal
speed up to 4 times higher than previous studies [1, 2] (figure A3, gray areas).
Similarly, the robot was capable of traversing a bump up to 3 hip height (figure 3(b), filled circles),
and also often did so dynamically at speeds higher than its walking-to-running transition speed decreasing
probability with bump height (figure 3(b), open circles). Despite our robot using open-loop control, its large
bump traversal probability was at least 30% higher than achieved in previous studies which robots used
carefully planned maneuvers [24 -- 26]. Although the portion of dynamic traversal among all traversal
decreased with bump height, it is worth noting that dynamic traversal was possible for all the bump height
tested that the animal or the robot was able to traverse (figure 3(c, d)).
9
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure 3. Probability of traversal (filled circles) and probability of dynamic traversal out of all trials (open
triangles) of the animal (a) and robot (b) as a function of bump height. The animal traversed dynamically
if its traversal speed was higher than its walking-to-running transition speed (see methods section 2.2 and
the companion paper [37]).
3.2. Locomotor modes and transitions
Although the bump obstacle was simple, the animal or robot displayed multiple locomotor modes
when attempting to traverse it (see section 2.2 for quantitative definitions of each mode):
(1) Climbing. The animal or robot often climbed against the large bump when its initial body yaw
and initial body yaw was small (i.e., body-terrain collision was head-on) (figure 4(a-f), supplementary
videos 1, 2). Climbing was most likely to result in successful traversal of the bump. When climbing resulted
in traversal, the animal or robot body pitched up to raise the head above the front face of the bump and
climbed over it while maintaining nearly 0° body yaw (figures 4(a, f), figures 4(c, d, e, f), red solid curves).
10
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
As bump height increased, the animal or robot body pitched up more to traverse (animal: up to 60°; robot:
up to 40°), but always recovered the initial orientation within 1 body length of forward displacement once
it was on the bump. When the animal climbed a high bump, it gripped the bump with its legs and pulled
itself up and forward onto the bump, while flexing the body ventrally up to 23° to allow its hind legs to
better brace against the front face of the bump (figure 4(a), frame 4). Although this appeared kinematically
similar to previous observations of quasi-static bump traversal [1, 2], in our study, the animal traversed a
bump up to twice as high and up to 8 times faster (figure S1, S6).
(2) Deflection. The animal or robot often was deflected against the large bump when its initial body
yaw and initial body yaw was large (figure 4(b, g), supplementary videos 1, 2). In this case, the body did
not pitch up substantially, nor did the head rise substantially (figure 4(c, h), dashed blue lines). Instead,
body yaw increased up to 90°, resulting in turning laterally and failure to traverse (figure 4(e, j), blue dashed
curves). Deflection most often resulted in failure to traverse.
It is worth emphasizing that for all but the smallest bump, both the animal and the robot had higher
initial head height, higher initial body pitch, and lower initial body yaw when traversing using climbing
than when being deflected (figure 4(c, d, e, h, I, j); animal: P < 0.05, repeated-measures ANOVA; robot: P
< 0.05, ANOVA). This is strong evidence that body-terrain interaction played a major role in large bump
traversal.
11
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure 4. Dynamic locomotion of the discoid cockroach and robot attempting to traverse a high bump.
Representative trials of (a, f) climbing traversal and (b, g) deflection. (c, h) Head height, zhead, as a function
of x-position of the head.* (d, i) Body pitch as a function of x-position of the head. (e, j) Body yaw as a
function of x-position of the head.† Solid red and blue dashed curves and shaded areas represent means ± 1
s.d. for the cases of climbing traversal and deflection. Data are shown for the 3 hip height (animal, 1.5 cm;
robot: 7.5 cm) bump as an example and have similar trends for other bump heights. For deflected trials,
* We noted that the head appeared to have slightly penetrated the bump after collision. This was due to the robot
head and body flexion after high-speed collision with the bump, which our head tracking method could not account
for (see methods in our companion paper for details [37]).
† Due to the ellipsoid-like shape of the animal body and the robot anterior, the head was only able to reach the bump
(xhead = 0) for body yaw < 45. As body yaw continued to increase during deflection, the head forward position xhead
actually decreased).
12
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
data were only shown until bump collision because the animal often moved backwards afterwards. Bracket
and asterisk represent statistically significant differences.
(3) Slanted climbing. Occasionally, the animal and robot climbed even with a high initial body yaw
and high initial body yaw, resulting in a slanted trajectory relative to the bump (figure 5(a), dark yellow
rectangle) as opposed to the trajectory perpendicular to the bump using head-on climbing. Careful
observation revealed that, during slanted climbing, the animal often transitioned between climbing and
deflection, with body yaw and velocity heading both oscillating by larger amplitudes that during head-on
climbing. This occurred more frequently as bump height increased (P < 0.0001, repeated-measures
ANOVA). To transition from the deflection to climbing, the animal often gripped and hung on to the front
face of the bump with a fore leg, pitching its body upwards and turning towards the bump with decreasing
body yaw. Once the second fore leg was able to reach the top of the bump, the leg and body kinematics
were similar to that during climbing. Because the robot operated with feed-forward control, its slanted
climbing was merely due to oscillations in body yaw and body pitch from intermittent leg-terrain
interaction. The robot did not actively transition between climbing and deflection like the animal did.
(4) Flipping over. The open-loop robot occasionally flipped over onto its dorsal side and failed to
traverse after impacting a large bump obstacle (3 and 4 hip height, figure 5(c, d), purple oval). The animal
was never flipped over in our experiments.
In addition, we found that the animal did not always use one locomotor mode when attempting to
traverse the large bump but frequently transitioned between modes, forming complex locomotor transition
pathways (figures 5). Unlike many previous studies [1, 3, 43, 44] where the focus was on a single locomotor
mode during obstacle traversal, such comprehensive observations of locomotor transition pathways [6, 18,
38, 45] provided a more realistic representation of animal locomotion in nature.
Comparison of locomotor transition pathways (figure 5) between each bump height revealed that,
as bump height increased, the animal or robot less frequently climbed and traversed, but were more often
13
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
deflected (animal: P < 0.0001; robot: P < 0.0001, multiple logistic regression). This change in the dominant
locomotor mode was the main reason why the traversal probability decreased with bump height.
14
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure 5. The animal and the robot used a variety of locomotor modes while attempting to traverse a large
bump. (a) Representative drawings of each locomotor mode. All trials begin with start (white diamond),
use one or a sequence of three modes -- climbing (red rectangle), deflection (blue rectangle), slanted
climbing (dark yellow rectangle), and end with traversal (white oval), flip over (purple oval), or failure to
traverse (exiting the field of view, not shown). The arrows represent the typical trajectory of the head for
each locomotor mode and the dashed curves show the projection of the trajectory to the surface. Animal (b,
c, d, e) and robot (f, g, h, i) locomotor pathways for each bump height (1, 2, 3, and 4 hip height,
respectively). Arrows indicate transition between modes. Arrow thickness is proportional to transition
probability, indicated by the number near each arrow.
3.3. Body-bump interaction is frequent
We observed that, besides legs, the animal's body frequently collided and continued to physically
interact with the bump during traversal, especially for larger bumps (figure 6). For the 2, 3, and 4 hip height
bump, collision probability was 34%, 61%, and 83%, respectively (figure 6, filled circles). This is in
contrast to previous studies on obstacle avoidance [5, 16, 17] or step climbing [1, 2, 6] where movement
was slow and obstacle collision was rare. The high probability of body-bump collision was because the
animal ran so fast that it did not have enough time to slow down even if antennae touched the bump. For
the majority of our trials (83%), the animal ran at speeds higher than the reaction time limit (figure A1, see
section 3.2 in our companion study [37]). Similarly, the open-loop robot also collided with the bump (figure
6, open circles) and continued to physically interacted with it frequently during traversal.
15
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure 6. The animal and the robot frequently collided with a large bump. Animal (filled circles) and robot
(open circles) probability of the body colliding with a bump increased with bump height. Error bars show
95% confidence interval.
It is worth emphasizing that these observations differed from previous studies of cockroach bump
traversal, where locomotion was much slower (up to 8 times) and the animal had sufficient time to use its
antenna to detect the bump, then slowed down or stopped and adjusted leg and body kinematics to traverse
or turn laterally, without body-bump collision [1, 19, 46].
3.4. Locomotion energy landscape model
Considering how frequent the animal or robot's body collided with and continued to physically
interact with the bump in our experiments, we speculated that body-bump interaction played an important
role in the observed sensitive dependence of traversal performance and locomotor transition pathways on
bump height. To understand this, we developed a simple locomotion energy landscape model [35] (figures
8, 9, supplementary video 3). The new concept of locomotion energy landscapes was recently introduced
to model body-terrain interaction during traversal of complex 3-D terrains such as grass-like beam
obstacles, where comprehensive contact mechanics models are still lacking [35]. Although locomotion
energy landscapes do not yet model leg-terrain contact or take into account the dynamics and non-
conservative forces of the system, they are useful in explaining how body-terrain interaction affect
locomotor transition pathways and traversal performance [35].
16
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure 7. Locomotion energy landscape model for 4 hip height bump traversal. (a, c, e) Body forward
position relative to the bump and body orientation. The ellipsoidal bodies show the body orientation prior
to bump encounter (black) and for climbing (red) and deflection (blue). (b, d, f) Locomotion energy
landscape. The energy landscape surface is a function of body yaw, body pitch, and x-position perpendicular
to the bump and is normalized by the body mass times bump height (E/(mgH)). The model is shown at two
representative locations: (a, b) just prior to the body colliding with the bump, xbody = −2/3 body length, (c,
d) during traversal, xbody = −1/3 body length. Red and blue circles indicate the animal's mean initial body
pitch and mean initial body yaw during pure climbing and pure deflection, respectively. Error bars represent
± s.d. Red and blue trajectories show the evolution of mean body pitch and mean body yaw as the animal
moved forward from xbody = −2/3 to xbody = −1/6 body length during pure climbing and pure deflection,
respectively. Brackets and asterisk represent a statistically significant difference.
We approximated the animal or robot body as a rigid ellipsoid with the same length, width, and
thickness as measured in experiments and a uniform mass distribution such that center of mass (CoM)
overlapped with geometric center. We assumed that the body had no overlap with the bump during traversal
17
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
and that its lowest point always touched the ground before encountering and after traversing the bump. In
addition, for simplicity, we approximated body roll as constantly zero because the animal or robot's body
roll oscillated around 0° by only a small magnitude (± 4°) and did not have a significant impact on traversal
(animal: P > 0.05, repeated-measures ANOVA; robot: P > 0.05, ANOVA). These constraints, together with
the invariance of the system in the lateral (y) direction, meant that the center of mass height, and thus the
gravitational potential energy of the body, depended only on its forward position relative to the bump, body
pitch, and body yaw, i.e., E = mgzCoM, and zCoM = zCoM(x, α, β).
To obtain the potential energy landscape resulting from body-bump interaction, we varied the
forward (+xbody) position of the body, and numerically calculated how potential energy depended on body
yaw, body pitch, and body roll (supplementary video 3). To understand how the locomotion energy
landscape changed as the animal or robot moved forward towards the bump during traversal (with
increasing xbody), we examined the landscape at two representative x positions (figure 7). For the following
discussions, we used results of a 4 hip height bump as an example. Modeling results of other bump heights
were qualitatively similar.
When the body was far away from the bump (xbody < -- 1/2 body length, figure 7(a, b)), its potential
energy only depended on body pitch (assuming constant zero body roll, see above). As the body moved
closer to the bump ( -- 1/2 body length < xbody < 0 body length, figure 7(c, d)), it had to turn laterally, pitch
up, and/or rise so as not to penetrate the bump. With no change or only a small change in body pitch or
body yaw, the body must rise over the edge of the bump so as not to penetrate it. This resulted in a large
"bump" in the potential energy landscape for small body pitch and body yaw (figure 7(d), black circle). As
body pitch and/or body yaw increased, the amount that the body had to rise decreased. This resulted in a
monotonic decrease of the potential energy as body pitch and body yaw initially increased from zero (e.g.,
figure 7(d), red and blue curves). For a given body yaw, as the body pitch continued to increase, the potential
energy reached a local minimum and then increased, resulting in a "valley" in the energy landscape. This
valley corresponded with all the body orientation states where the body was in contact with the ground and
18
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
the top edge or front face of the bump at the same time. The valley became deeper as body yaw increased
and eventually reached a potential energy well (global minimum) (figure S2).
3.5. Model explained dependence of bump traversal on body-terrain interaction
A quasi-static stability analysis using the topology of the potential energy landscape provided
insight into how initial body pitch and initial body yaw affected locomotor modes for a given bump height.
The potential energy landscape during body-bump contact had a saddle point, i.e., the local minimum in
the body pitch direction at a finite positive body pitch and zero body yaw (figure 7(d)). As the body moved
closer and came into contact with the bump, a pitched-up body was most stable along the pitch direction
(stable equilibrium in the pitch direction), but this pitched-up body orientation was unstable in the yaw
direction (unstable equilibrium). This suggested that, except when the animal or robot always approached
the bump head-on (maintaining zero body yaw), it tended to be deflected to either side (increasing body
yaw) as long as there was any slight body yaw oscillation (figure 7(d), blue curve). In addition, when the
body was deflected (increasing body yaw), it became attracted towards the potential energy well along the
valley where it was eventually trapped in an orientation with high body yaw and low body pitch (bottom of
figure 7(d)). During slanted climbing, the animal had to overcome this tendency to be deflected further.
Finally, the model also provided insight into how bump height affected body-terrain interaction and thus
traversal performance. As bump height increased, we found that the potential energy well became deeper
and more difficult to escape (figure S2).
These stability analyses revealed two general principles for large bump traversal during which
body-terrain interaction played a major role. First, a head-on approach (small initial body yaw) and a pitched
up body posture facilitate traversal using climbing. Indeed, we observed that for all except the smallest
bump height, both the animal and the robot had a low initial body yaw and a high initial body pitch when
they traversed the bump via climbing (figure 4(c, d, e, h, i, j), figure S4(b, c, d, e)). By contrast, when they
were deflected and failed to traverse, initial body yaw was significantly higher (animal: P < 0.0001; robot:
P < 0.0001, multiple logistic regression). Second, as bump height increases, traversal via climbing becomes
19
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
less likely, because it is more difficult to escape the deeper potential energy well that the body can get
deflected into. Indeed, we observed both the animal and robot were more often deflected with lower
traversal probability as bump height increased (animal: P < 0.0001; robot: P < 0.0001, multiple logistic
regression).
Finally, the model also suggested that a pitched-up body posture, as opposed to a horizontal posture
(zero body pitch), also facilitated climbing and traversal. Examination of the potential energy landscape as
a function of body pitch and body forward position (figure S3) showed that, when initial body pitch was
near zero, the body had to climb up a steep, high potential barrier over a very small forward displacement.
Just a slight increase of body pitch reduced the slope of this potential energy barrier, making traversal easier.
We noted that all these analyses were using a simple quasi-static body-terrain interaction model
and neglected body and leg dynamics. In reality, the animal or robot ran into the bump with high speed and
pushed its legs against the ground during traversal. Ground reaction forces, together with high body
momentum (if properly re-directed), could also played a role in traversal of large bump obstacles during
dynamic locomotion, and should be better understood in future work.
3.6. Tail-assisted pitch control to enhance high bump traversal
The experimental observations that successful traversal had significantly higher initial body pitch
(figure 4(d, i), figure 7(b)) and higher initial head height (figure 4(c, h)) and the above model insights that
increasing body pitch facilitated climbing inspired us to propose a novel control strategy to use active body
pitching to increase dynamic bump traversal performance of the robot. To demonstrate this, we tested the
tailed robot developed in our companion study [37].
While running at a constant running speed of 70 ± 3 cm s−1, the robot's active tail increased its
initial body pitch from 0° ± 4° to 3° ± 7° (P = 0.04, ANOVA). This small increase in initial body pitch not
only increased traversal probability for all except the smallest bump tested, but also increased the maximal
traversable bump height for the tailed robot from 2 hip height to 3.5 hip height, a 75% increase (figure 8).
20
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Although the increase in initial body pitch by the active tail appeared small, it raised the head by 1.3 cm,
about 15% of the height of the highest bump traversable by the robot (3.5 hip height). This likely increased
the probability of the robot's head to reach over the front corner of the bump (figure S3).
We noted that the timing of tail actuation was important for pitching up the body at the right
moment so that the head reached over the top of the bump, particularly for the larger 3 and 4 hip height
bumps. Pitching up too late did not allow the head to clear the bump in time, while pitching up too early
did not work because the increase of body pitch using an active tail was temporary. Future studies should
add fast sensors to detect impending bump obstacles during rapid locomotion and precisely control the
magnitude and timing of active body pitching [47 -- 49] to further increase traversal performance.
Figure 8. Bump traversal probability for the tailed robot, with (open circles) and without (filled circles) tail
actuation as a function of bump height.
3.7. The role of sensory feedback in traversal
Although the rapid-running animal frequently collided with the high bump and body-terrain
interaction strongly affected whether it climbed to traverse or was deflected, the animal likely used sensory
feedback to actively adjusted its body and leg kinematics for traversal, as well as to overcome less
advantageous body-bump interaction. This was evidenced by a few observations. First, during climbing,
especially for larger bump height, the animal often actively flexed its neck joint and abdomen and used its
21
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
legs to grip and pull itself onto the bump. Second, during slanted climbing, although the animal was
susceptible to being deflected given its large initial body yaw, it only occasionally lost grip and eventually
become deflected (6% and 12% for the 3 and 4 hip height bump, respectively, figure 5(d, e, h, i)). Third,
the animal was still able to traverse a 3 hip height bump even when initial body yaw was up to 60° for the
3 hip height bump and up to 50° for the 4 hip height bump.
By contrast, the sensor-less, open-loop robot had a poorer ability to grip with its legs and lacked
body flexibility and had no ability to actively adjust its body orientation or leg kinematics. Compared to the
animal, it more often failed by falling backward during climbing (figure 5(h, i)). In addition, the robot was
unable to traverse a 3 hip height bump when initial body yaw exceeded 30° and was never able to traverse
the 4 hip height bump. Further, the robot occasionally lost stability during climbing and flipped over, which
was never seen in the animal experiment. All these limitations led the robot to have worse traversal
performance than animal, particularly for the large bump height figure 5(e, i). We expect that adding
controlled body flexion guided by the locomotion energy landscape model could further increase maximal
traversable bump height for robots -- a previous study had doubled this limit using empirically tuned body
flexion [23, 29, 50]. Future studies should also add adhesive mechanisms [51 -- 53] to improve the robot's
gripping ability.
Finally, previous studies have shown that whether cockroaches can locate the top of a bump with
its antennae determines whether it climbs over or under a shelf during slow locomotion [19]. When the
animal's body was more pitched up with its head higher and redirected upwards during collision, it might
have been better able to locate the top of the bump with its antennae and further make kinematic adjustments
to climb [1, 2]. By contrast, when the animal's initial body pitch was low and its head deflected laterally, it
might not have been able to do so and instead perceived the bump as a wall-like obstacle and continued to
turn and follow the wall [54, 55].
3.8. Similarity between large bump and gap traversal
22
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Comparing the current study with our companion study [37], we found that, surprisingly, dynamic
traversal of large gap and bump obstacles share a few common features. First, it was possible to dynamically
traverse both a large gap and a high bump comparable to body size (and much larger than observed in
previous studies [1, 2, 19, 42]), even for the open-loop robot. In addition, during rapid dynamic locomotion
over such a large obstacle, because body-terrain contact was nearly un-avoidable and sensor-based
adjustment and planning became unfeasible, body-terrain interaction had a strong impact on traversal.
Despite the distinct differences between a large gap and a high bump, traversal of both obstacles was
facilitated by rapid running towards the obstacle with a pitched-up, head-on body orientation. Finally, for
both obstacles, the animal out-performed the robot because it was able to better grip and brace itself using
sensory feedback, and active body pitching increased the robot's maximal traversal performance.
4. Conclusions
In this study, we discovered that small insects and legged robots are capable of dynamically
traversing a bump obstacle much higher than its hip height. Remarkably, the discoid cockroach was capable
of dynamically traversing a large bump up to 4 times its hip height at speeds comparable to [41, 56] or
higher [1, 2, 57 -- 61] than during running, and our open-loop robot was capable of dynamically traversing a
high bump up to 3 times its hip height at speeds above the walking-to-running transition speed. These newly
discovered performance limits for dynamic bump traversal were more than 200% that of previous animal
studies [1, 2] and more than 30% higher than previous robot studies [24 -- 26]. In addition, we discovered
that body-terrain interaction strongly affected locomotor modes and traversal performance. Furthermore,
our locomotion energy landscape model explained why traversal was less likely as bump height increased
and why a head-on and slightly pitched-up body orientation facilitated dynamic bump traversal by reducing
the likelihood of being deflected. These experimental and modeling insights allowed us to use active body
pitching to increase maximal traversable bump height by 75%. Finally, the animal's traversal performance
23
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
exceeded that of the robot thanks to its ability to use sensory feedback to make body and leg adjustments
and its better gripping ability and body flexibility.
This study also expanded the novel approach of locomotion energy landscapes [35] in modeling
body-terrain interaction to large bump obstacles, another representative 3-D terrain. Besides sensor-based
planning and decision processes using geometric information of the environment [19, 46, 62, 63], physical
interaction of animal/robot body with the terrain also played an important role in initiating locomotor
transitions and determining performance during large obstacle traversal. Together with recent success of
this approach in understanding how locomotor shape affect traversal of cluttered grass-like obstacles [35],
our study is further establishing locomotion energy landscapes as a general framework for modeling
locomotor-terrain interaction in a diversity of complex 3-D terrains.
Together, our two companion studies ([37] and this paper) are a major step towards establishing
the new field of terradynamics of animal and robot locomotion in complex terrains. First, the remarkable
performance limits of dynamic traversal of large gap and bump obstacles from our studies provided new
insights into how well and by what means animals can move in the complex natural world. The ability to
dynamically traverse large gaps and bumps comparable to body size during rapid predator-prey pursuit is
crucial for the survival of insects and many other legged animals that live on forest floor [64], in scrubs [9]
and crop fields [65, 66], and in rocky environments [13]. Our studies provide the first mechanistic
explanations of how the need to survive exerts selective pressure on the ability of both predators and prey
to run fast and maintain appropriate body posture so that traversal is likely and without significant stumble
or slow down, even when sensing becomes unfeasible.
Second, we expect that the performance limits, mechanical principles, and novel control strategies
revealed by our two companion studies will inspire a variety of robots to better take advantage of kinetic
energy and momentum of rapid running and simple body posture control to traverse large gap- and bump-
like obstacles. This will not only add high-speed dynamical traversal of large gap and high bump obstacles
24
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
to the locomotor repertoire of legged robots, but also simplify their lower-level control so that sensing and
computation resources can be better devoted to deliberate maneuvers and precise planning [24 -- 26, 30]
required for truly large un-traversable obstacles beyond the limits by using dynamics. Such advancements
will expand the operation time and accessible terrain area for robots that perform tasks in complex 3-D
terrains such as building rubble and landslides.
Finally, considering their distinct differences in geometry, it is surprising that similar principles
exist for traversal of both a large bump and a large gap during rapid locomotion. This suggests that there
may be general principles of dynamic locomotion in a diversity of complex 3-D terrains. Future
terradynamic studies should further explore the locomotor and terrain parameter space to reveal these
general principles and enable quantitative understanding and predictions of the movement of terrestrial
animals and robots.
Acknowledgements
We thank Yuanfeng Han, Tom Libby, Simon Sponberg, Bob Full, and two anonymous reviewers
for helpful discussions; Nastasia Winey and Rafael de la Tijera Obert for help with preliminary
experiments; and Yuanfeng Han for help with experimental setup. This work is funded by a Burroughs
Wellcome Fund Career Award at the Scientific Interface, an Army Research Office Young Investigator
Award, and The Johns Hopkins University Whiting School of Engineering start-up funds to C L. Author
contributions: S W G designed study, performed experiments, analyzed data, developed model, and wrote
the paper; C L designed and supervised study, assisted with model development, and wrote the paper.
References
[1]
Watson J T, Ritzmann R E, Zill S N and Pollack A J 2002 Control of obstacle climbing in the
cockroach, Blaberus discoidalis. I. Kinematics J. Comp. Physiol. A 188 39 -- 53
25
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Watson J T, Ritzmann R E and Pollack A J 2002 Control of climbing behavior in the cockroach,
Blaberus discoidalis. II. Motor activities associated with joint movement J. Comp. Physiol. A 188
55 -- 69
Zurek D B and Gilbert C 2014 Static antennae act as locomotory guides that compensate for
visual motion blur in a diurnal , keen-eyed predator Proc. R. Soc. B 281 20133072
Lewinger W a., Harley C M, Watson M S, Branicky M S, Ritzmann R E and Quinn R D 2009
Animal-inspired sensing for autonomously climbing or avoiding obstacles Appl. Bionics Biomech.
6 43 -- 61
Baba Y, Tsukada A and Comer C M 2010 Collision avoidance by running insects: antennal
guidance in cockroaches. J. Exp. Biol. 213 2294 -- 302
Kohlsdorf T and Biewener A A 2006 Negotiating obstacles: Running kinematics of the lizard
Sceloporus malachiticus J. Zool. 270 359 -- 71
Tucker D B and Mcbrayer L D 2012 Overcoming obstacles: The effect of obstacles on locomotor
performance and behaviour Biol. J. Linn. Soc. 107 813 -- 23
Olberding J P, McBrayer L D and Higham T E 2012 Performance and three-dimensional
kinematics of bipedal lizards during obstacle negotiation J. Exp. Biol. 215 247 -- 55
Parker S E and McBrayer L D 2016 The effects of multiple obstacles on the locomotor behavior
and performance of a terrestrial lizard J. Exp. Biol. 219 1004 -- 13
Birn-Jeffery A V, Hubicki C M, Blum Y, Renjewski D, Hurst J W and Daley M A 2014 Don't
break a leg: running birds from quail to ostrich prioritise leg safety and economy on uneven
terrain. J. Exp. Biol. 217 3786 -- 96
Williams N and Pearson K G 2007 Stepping of the forelegs over obstacles establishes long-
lasting memories in cats Curr. Biol. 17 R621-623
Stierle I E, Getman M and Comer C M 1994 Multisensory control of escape in the cockroach
Periplaneta americana I. Initial evidence from patterns of wind-evoked behavior J. Comp. Physiol.
A 174 1 -- 11
Goodman B A 2007 Divergent morphologies , performance , and escape behaviour in two tropical
rock-using lizards ( Reptilia : Scincidae ) 85 -- 98
Murphy R R, Tadokoro S, Nardi D, Jacoff A, Fiorini P, Choset H and Erkmen A M 2008 Search
and Rescue Robotics Handbook of Robotics (Springer) pp 1151 -- 73
[15]
Pratt G A 2014 Robot to the rescue Bull. At. Sci. 70 63 -- 9
Blaesing B and Cruse H 2004 Stick insect locomotion in a complex environment: climbing over
large gaps J. Exp. Biol. 207 1273 -- 86
Theunissen L M, Vikram S and Dürr V 2014 Spatial coordination of foot contacts in unrestrained
climbing insects. J. Exp. Biol. 217 3242 -- 53
Theunissen L M and Durr V 2013 Insects use two distinct classes of steps during unrestrained
locomotion PLoS One 8 1 -- 18
Harley C M, English B a and Ritzmann R E 2009 Characterization of obstacle negotiation
behaviors in the cockroach, Blaberus discoidalis. J. Exp. Biol. 212 1463 -- 76
26
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Daley M A, Usherwood J R, And G F and Biewener A A 2006 Running over rough terrain:
guinea fowl maintain dynamic stability despite a large unexpected change in substrate height J.
Exp. Biol. 209
Birn-Jeffery a. V. and Daley M 2012 Birds achieve high robustness in uneven terrain through
active control of landing conditions J. Exp. Biol. 215 2117 -- 27
Morrey J M, Lambrecht B, Horchler a. D, Ritzmann R E and Quinn R D 2003 Highly mobile and
robust small quadruped robots Proc. 2003 IEEE/RSJ Int. Conf. Intell. Robot. Syst. (IROS 2003)
(Cat. No.03CH37453) 1 0 -- 5
Wei T E, Quinn R D and Ritzmann R E 2005 Robot designed for walking and climbing based on
abstracted cockroach locomotion mechanisms Proceedings, 2005 IEEE/ASME Int. Conf. Adv.
Intell. Mechatronics. 24 -- 8
Chou Y-C, Yu W-S, Huang K-J and Lin P-C 2012 Bio-inspired step-climbing in a hexapod robot
Bioinspir. Biomim. 7 36008
Johnson A M and Koditschek D E 2013 Toward a vocabulary of legged leaping Proc. - IEEE Int.
Conf. Robot. Autom. 2568 -- 75
Chou Y C, Huang K J, Yu W S and Lin P C 2015 Model-based development of leaping in a
hexapod robot IEEE Trans. Robot. 31 40 -- 54
Mondada F, Pettinaro G C, Guignard A, Ivo W, Floreano D, Deneubourg J and Nolfi S 2004
Swarm-Bot : a New Distributed Robotic Concept Auton. Robots 17 1 -- 40
Raibert M, Blankespoor K, Nelson G and Playter R 2008 BigDog , the Rough-Terrain Quadruped
Robot vol 41(IFAC)
Boxerbaum A S, Oro J, Peterson G and Quinn R D 2008 The latest generation Whegs robot
features a passive-compliant body joint IEEE/RSJ Int. Conf. Intell. Robot. Syst. 1636 -- 41
Brill A L, De A, Johnson A M and Koditschek D E 2015 Tail-assisted rigid and compliant legged
leaping IEEE Int. Conf. Intell. Robot. Syst. 6304 -- 11
More H L, Hutchinson J R, Collins D F, Weber D J, Aung S K H and Donelan J M 2010 Scaling
of sensorimotor control in terrestrial mammals. Proc. Biol. Sci. 277 3563 -- 8
Mongeau J-M, Sponberg S N, Miller J P and Full R J 2015 Sensory processing within cockroach
antenna enables rapid implementation of feedback control for high-speed running maneuvers. J.
Exp. Biol. 218 2344 -- 54
Full R J and Koditschek D E 1999 Templates and anchors: neuromechanical hypotheses of legged
locomotion on land J. Exp. Biol. 2 3 -- 125
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.
Bioinspir. Biomim. 2 9 -- 18
Li C, Pullin A O, Haldane D W, Lam H K, Fearing R S and Full R J 2015 Terradynamically
streamlined shapes in animals and robots enhance traversability through densely cluttered terrain
Bioinspir. Biomim 10
[36]
Okada J and Toh Y 2000 The role of antennal hair plates in object-guided tactile orientation of the
cockroach (Periplaneta americana) J. Comp. Physiol. A 186 849 -- 57
27
[37]
[38]
[39]
[40]
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Gart S, Othoyoth R, Ren Z, Yan C and Li C 2017 Dynamic locomotion of insects and legged
robots over large obstacles I. Body dynamics and terrain interaction reveal a template for dynamic
gap traversal Bioinspir. Biomim
Han Y, Luo Y, Bi J and Li C 2017 Body shape affects yaw and pitch motions of insects traversing
complex 3-D terrains Integr. Comp. Biol. 168
Crall J D, Gravish N, Mountcastle A M and Combes S A 2015 BEEtag : A Low-Cost , Image-
Based Tracking System for the Study of Animal Behavior and Locomotion PLoS One 10
e0136487
Daltorio K a., Tietz B R, Bender J a., Webster V a., Szczecinski N S, Branicky M S, Ritzmann R
E and Quinn R D 2013 A model of exploration and goal-searching in the cockroach, Blaberus
discoidalis Adapt. Behav. 21 404 -- 20
[41]
Full B Y R J and Tu M S 1990 Mechanics of six-legged runners 148 129 -- 46
[42]
[43]
[44]
[45]
[46]
[47]
[48]
[49]
[50]
[51]
[52]
[53]
Sponberg S and Full R J 2008 Neuromechanical response of musculo-skeletal structures in
cockroaches during rapid running on rough terrain. J. Exp. Biol. 211 433 -- 46
Ritzmann R E, Pollack A J, Archinal J, Ridgel A L and Quinn R D 2005 Descending control of
body attitude in the cockroach Blaberus discoidalis and its role in incline climbing J. Comp.
Physiol. A 191 253 -- 64
Goldman D I, Chen T S, Dudek D M and Full R J 2006 Dynamics of rapid vertical climbing in
cockroaches reveals a template. J. Exp. Biol. 209 2990 -- 3000
Daley M and Biewener A A 2006 Running over rough terrain reveals limb control for intrinsic
stability. Proc. Natl. Acad. Sci. U. S. A. 103 15681 -- 6
Okada J and Toh Y 2006 Active tactile sensing for localization of objects by the cockroach
antenna J. Comp. Physiol. A 192 715 -- 26
Qian F and Goldman D 2015 Anticipatory control using substrate manipulation enables trajectory
control of legged locomotion on heterogeneous granular media Zebra-tailed Proc. SPIE 9467 1 -- 12
Jusufi a, Kawano D T, Libby T and Full R J 2010 Righting and turning in mid-air using
appendage inertia: reptile tails, analytical models and bio-inspired robots. Bioinspir. Biomim. 5
45001
Libby T, Moore T Y, Chang-Siu E, Li D, Cohen D J, Jusufi A and Full R J 2012 Tail-assisted
pitch control in lizards, robots and dinosaurs Nature 481 181 -- 4
Casarez C S and Fearing R S 2016 Step Climbing Cooperation Primitives for Legged Robots with
a Reversible Connection IEEE Int. Conf. Robot. Autom. 3791 -- 8
Bibuli M, Caccia M and Lapierre L 2008 Biologically Inspired Climbing with a Hexapedal Robot
J. F. Robot. 25 223 -- 42
Gorb S N 2002 Structural Design and Biomechanics of Friction-Based Releasable Attachment
Devices in Insects Integr. Comp. Biol. 42 1127 -- 39
Daltorio K A, Wei T E, Horchler A D, Southard L, Wile G D, Quinn R D, Gorb S N and
Ritzmann R E 2009 Mini-Whegs TM Climbs Attachment Mechanisms Int. J. Rob. Res. 28 285 --
302
[54]
Cowan N J and Full R J 2006 Task-level control of rapid wall following in the American
28
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
cockroach J. Exp. Biol. 209 3043 -- 3043
Daltorio K A, Mirletz B T, Sterenstein A, Cheng J C, Watson A, Kesavan M, Bender J A,
Ritzmann R E and Quinn R D 2014 How Cockroaches Employ Wall-Following for Exploration
72 -- 83
Ting L H, Blickhan R and Full R J 1994 Dynamic and Static Stability in Hexapedal Runners J.
Exp. Biol. 197 251 -- 69
Ritzmann J T W R E 1998 Leg kinematics and muscle activity during treadmill running in the
cockroach , Blaberus discoidalis : II . Fast running J. Comp. Physiol. A 182 23 -- 33
Ritzmann J T W R E 1998 Leg kinematics and muscle activity during treadmill running in the
cockroach , Blaberus discoidalis : I . Slow running J. Comp. Physiol. A 182 11 -- 22
Tryba A K and Ritzmann R O Y E 2000 Multi-Joint Coordination During Walking and Foothold
Searching in the Blaberus Cockroach . II . Extensor Motor Neuron Pattern J. Neurophysiol. 83
3337 -- 50
Tryba A K and Ritzmann R O Y E 2000 Multi-Joint Coordination During Walking and Foothold
Searching in the Blaberus Cockroach . I . Kinematics and Electromyograms J. Neurophysiol. 83
3323 -- 36
Ridgel A L and Ritzmann R E 2005 Effects of neck and circumoesophageal connective lesions on
posture and locomotion in the cockroach J. Comp. Physiol. A 191 559 -- 73
Leonard J J and Durrant-Whyte H F 1991 Simultaneous map building and localization for an
autonomous mobile robot Intell. Robot. Syst. 3 1442 -- 7
Thrun S, Burgard W and Fox D 2000 A real-time algorithm for mobile robot mapping with
applications to\nmulti-robot and 3D mapping IEEE Int. Conf. Robot. Autom. 1 321 -- 8
Bell W J, Roth L M and Nalepa C A 2007 Cockroaches. Ecology, Behaviour and Natural History
(Baltimore, MD, USA: The Johns Hopkins University Press)
Michalková V and Pekár S 2009 How glyphosate altered the behaviour of agrobiont spiders (
Araneae : Lycosidae ) and beetles ( Coleoptera : Carabidae ) Biol. Control 51 444 -- 9
Griffiths E, Wratten S and Vickerman G 1985 Foraging by the carabid Agonum dorsde in the
field Ecol. Entomol. 10 181 -- 9
29
[55]
[56]
[57]
[58]
[59]
[60]
[61]
[62]
[63]
[64]
[65]
[66]
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Supplementary Figures
Figure S1. Histogram of approach speeds for animal bump experiments. Vertical dashed line represents an
approach speed of 30 cm s−1, Fr = 1.5, above which value the animals are unlikely to react obstacle
encounter (see section 3.3).
30
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S2. Energy well depth as a function of forward position of the body for each bump height. The
energy well depth for each forward position of the body was calculated by taking the difference between
the potential energy of a body with zero body pitch and body yaw (figure 9(b, d, f), black circle) and the
minimal potential energy on the landscape (for example, figure 9(b), 0° body pitch).
31
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S3. Locomotion energy landscape as a function of body pitch and forward position of the body
(shown for the 4 hip height bump as an example). Potential energy is normalized by body mass and bump
height. The white arrows show the potential energy of the body at representative initial body pitch angles
as it translates towards the bump. The white dashed curve on the surface shows where the body contacts
the bump. The red arrow shows the range of body pitch where the head reaches over the near top corner of
the bump.
32
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S4. Approach kinematics as function of bump height for the animal (a, b, c) and robot (d, e, f)
experiments. (a, d) Initial body pitch as function of bump height. (b, e) Approach speed as function of bump
height. (c, f) Initial body yaw as function of bump height. Red filled circles represent climbing traversals
and blue open circles represent deflections. Error bars represent ± 1 s.d. Asterisks in indicate a statistically
significance difference between the cases of climbing and deflection (P < 0.05, multiple logistic regression).
33
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S5. Histogram of initial body yaw for trials that traverse the bump and trials that deflect and fail to
traverse for animals (a) and robot (b).
34
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S6. Traversal speed as a function of approach speed. The dashed lines show the animal's walking-
to-running transition speed. Asterisks represent significant non-zero slopes (linear least squares regression).
†Shows the of approach speed for previous studies of cockroach bump traversal [1], [2].
35
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S7. Dynamic locomotion of the animal (a, b, c, d) and robot (e, f, g, h) over a 1 hip height bump
obstacle. (a, e) Head height as a function of forward of the head. (b, f) Body pitch as a function of forward
of the head. (c, g) Body roll as a function of forward of the head. (d, h) Body yaw as a function of forward
of the head. Solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of
climbing and deflection. For simplicity, only movement perpendicular to the bump (within the x-z plane) is
shown.
36
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S8. Dynamic locomotion of the animal (a, b, c, d) and robot (e, f, g, h) over a 2 hip height bump
obstacle. (a, e) Head height as a function of forward of the head.* (b, f) Body pitch as a function of forward
of the head. (c, g) Body roll as a function of forward of the head. (d, h) Body yaw as a function of forward
of the head. Solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of
climbing and deflection. For simplicity, only movement perpendicular to the bump (within the x-z plane) is
shown.
* The slight "penetration" of the head into the bump in (a, d) was an artifact due to bending of the neck joint of the
animal or deformation of the robot anterior during collision).
37
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Figure S9. Dynamic locomotion of the animal (a, b, c, d) and robot (e, f, g, h) over a 4 hip height bump
obstacle. (a, e) Head height as a function of forward of the head.* (b, f) Body pitch as a function of forward
of the head. (c, g) Body roll as a function of forward of the head. (d, h) Body yaw as a function of forward
of the head. Solid red and blue dashed curves and shaded areas represent means ± 1 s.d. for the cases of
climbing and deflection. For simplicity, only movement perpendicular to the bump (within the x-z plane) is
shown.
* The slight "penetration" of the head into the bump in (a, d) was an artifact due to bending of the neck joint of the
animal or deformation of the robot anterior during collision).
38
Bioinspiration & Biomimetics (2018), 13, 026005; https://li.me.jhu.edu
Supplementary Video 1
https://www.youtube.com/watch?v=7ZfMls2l5O0
Supplementary Video 2
https://www.youtube.com/watch?v=_I5vM6PgSbY
Supplementary Video 3
https://www.youtube.com/watch?v=v1vXGOw5I_Q
Supplementary Video 4
https://www.youtube.com/watch?v=ouPlzMfjN4k
39
|
1811.09548 | 1 | 1811 | 2018-11-23T16:37:41 | The Range of Cooperativity Modulates Actin Binding Protein Cluster Size, Density and Dynamics | [
"physics.bio-ph",
"q-bio.SC"
] | The actin cytoskeleton is composed of multiple networks which are specialized for several processes such as cell motility or cell division. Each of these networks are composed of organized actin microfilaments which are decorated with specific sets of actin binding proteins (ABPs). The molecular mechanisms guiding ABPs to specific actin networks are still poorly understood, but cooperativity, the mechanism by which the binding of an ABP is positively influenced by proximal bound ABPs, plays a crucial role in generating locally dense stretches of ABPs. Cooperative binding is characterized by its amplitude, but also by the range at which its effects are propagated along an actin filament through long-range allosteric interactions. The range of these allosteric effects is still debated, but is likely to be significant at the lengthscale of actin filaments in cells. Here, we investigated how cooperativity influences the clustering of ABPs, using a stochastic computational model of binding of ABPs to actin filaments. The model reproduces the formation of ABP clusters observed experimentally at the single filament scale, and provides a theoretical estimation of the range of cooperativity for proteins such as ADF/cofilin. We found that both the amplitude and the spatial range of cooperativity dramatically impact the properties of clustering. However, the parameters of cooperativity modulate differently the rate of assembly, size and dynamics of the ABP clusters, suggesting that cooperativity is an efficient mechanism to regulate precisely the recruitment of ABPs in cells. This work provides a more general framework for future understanding of how actin networks acquire distinct and specific protein compositions from a common cytoplasm. | physics.bio-ph | physics |
The Range of Cooperativity Modulates Actin Binding Protein
Cluster Size, Density and Dynamics
Thomas Le Goff1 and Alphée Michelot1
1Aix Marseille Univ, CNRS, IBDM, Turing Centre for Living Systems, Marseille, France
Abstract
The actin cytoskeleton is composed of multiple networks which are specialized for several pro-
cesses such as cell motility or cell division. Each of these networks are composed of organized
actin microfilaments which are decorated with specific sets of actin binding proteins (ABPs). The
molecular mechanisms guiding ABPs to specific actin networks are still poorly understood, but co-
operativity, the mechanism by which the binding of an ABP is positively influenced by proximal
bound ABPs, plays a crucial role in generating locally dense stretches of ABPs. Cooperative binding
is characterized by its amplitude, but also by the range at which its effects are propagated along
an actin filament through long-range allosteric interactions. The range of these allosteric effects is
still debated, but is likely to be significant at the lengthscale of actin filaments in cells. Here, we
investigated how cooperativity influences the clustering of ABPs, using a stochastic computational
model of binding of ABPs to actin filaments. The model reproduces the formation of ABP clusters
observed experimentally at the single filament scale, and provides a theoretical estimation of the
range of cooperativity for proteins such as ADF/cofilin. We found that both the amplitude and
the spatial range of cooperativity dramatically impact the properties of clustering. However, the
parameters of cooperativity modulate differently the rate of assembly, size and dynamics of the ABP
clusters, suggesting that cooperativity is an efficient mechanism to regulate precisely the recruitment
of ABPs in cells. This work provides a more general framework for future understanding of how
actin networks acquire distinct and specific protein compositions from a common cytoplasm.
Introduction
The actin cytoskeleton is an ensemble of biopolymers that cells use to power diverse functions such
as cell motility or cell division. Actin filaments are not randomly organized, but assemble in a vari-
ety of structures whose properties are optimized for a given cellular function [35]. In particular, the
organization and dynamics of the filaments is tightly controlled in time and space by specific sets of
Actin Binding Proteins (ABPs). These ABPs bind to the side or to the ends of the filaments, and
regulate important biochemical reactions such as their elongation rates, their crosslinking or their rate
of disassembly [1, 31].
Importantly, individual actin networks are specifically regulated because they interact with defined
sets of ABPs. While they assemble, actin networks must precisely control the accessibility of the
individual filaments to the appropriate families of ABPs [23, 27]. As all the actin filaments of the cell
are built from identical actin subunits, cells must employ efficient molecular mechanisms to specify in
time and space the identity of actin filaments. As a matter of fact, if ABPs were systematically following
the law of mass action, they would decorate with a similar ratio all the actin filaments of the cell. In this
study, we focused on cooperativity, which for actin filaments is the property that the binding of a first
ABP influences positively the attachment of other ABPs by increasing locally the association constant
and/or decreasing their dissociation constant. Cooperativity represents an efficient mechanism to favor
a high local density of identical ABPs, even if this family of proteins may represent only a small fraction
of all the ABPs present in the cytoplasm.
The cooperative binding of proteins to the side of actin filaments can occur through two non-exclusive
mechanisms. The first mechanism of cooperativity derives from possible interactions between two ABPs
bound to consecutive sites of an actin filament. An affinity between neighboring ABPs increases the
likelihood that an ABP will bind next to a previously bound one, inducing a cooperative effect. This
1
mechanism, also called end-to-end cooperativity, is the main mechanism of cooperativity for proteins
such as tropomyosins [39, 5]. Tropomyosins are multi-functional coiled coil dimers that wrap around
actin filaments. They bind adjacently, without gaps, because the N-terminus of one tropomyosin sub-
unit can directly bind to the C-terminus of another tropomyosin subunit along the actin filament [14].
As tropomyosin extend over several subunits of actin, end-to-end cooperativity represents an efficient
mechanism to limit the existence of empty gaps on the actin filament between consecutive tropomyosins.
The second mechanism of cooperativity is independent of any potential side interaction between
ABPs. Rather, a body of literature documents how actin binding proteins may also affect structurally
their substrate, the actin filament [10, 33, 27, 21]. Importantly, all the subunits of an actin filament
do not adopt structural conformations independently from one another, but rather adopt preferential
conformations over several actin subunits through long-range allosteric interactions [30, 24, 11, 17]. Thus,
ABPs can modify locally the structure of actin filaments, which in turn favors the binding of additional
ABPs and trigger cooperativity. This mechanism of cooperativity implies that cooperative binding does
not only occur strictly on adjacent sites along actin filaments, but rather stochastically over the distance
where the structure of the actin filament has been modified. Such mechanism of cooperativity applies
for proteins such as ADF/cofilin, which are small globular proteins implicated in actin disassembly.
The binding of ADF/cofilin to actin is equimolar, and induces major conformational changes on actin
filaments [10, 4, 38]. Principally, ADF/cofilin changes the mean twist of actin filaments and increases
their flexibility [25, 9, 7]. The cooperativity of ADF/cofilin binding to actin filaments has been reported
extensively, although attempts to evaluate a range of cooperativity for ADF/cofilin along actin filaments
brought considerably different results among studies. Differential scanning calorimetry experiments and
spectroscopic lifetime measurements provided originally an order of magnitude of about 100 subunits
for the structural changes induced by ADF/cofilin along an actin filament [3, 32]. Direct observations
include (1) real-time fluorescence microscopy, which evaluates a range of cooperative binding of about
24 actin subunits [15]; (2) atomic force microscopy, which revealed asymmetric conformational changes
in filaments induced by ADF/cofilin propagated over one-half of a helix (i.e. about 15 subunits) [29];
and (3) cryo-EM data which did not find any visible modification of the twist of actin filaments further
than the first few subunits from the boundary between bare and decorated segments of the filament [20].
The uncertainty in the range of allosteric modifications of the actin filaments is problematic for our
interpretation of these effects, as actin filaments are on average very short in cells. For example, in the
branched network of the lamellipodium, the spacing between Y-junctions occurs within 20-50 nm, which
corresponds to 8-18 subunits of actin [36]. At endocytic sites, the spacing between Y-junctions occurs
within 50-200 nm, which corresponds to 18-68 subunits of actin [44, 34]. For linear networks, actin
filaments are in general longer, for example in filopodia [37], or in the cytokinetic ring with an average
length of 0.8 µm, which corresponds to 270 actin subunits [22]. Nevertheless, one understands easily
that a range of cooperativity of 1 subunit or 100 subunits for actin filaments in cells will have major
consequences for their decoration by ABPs. Moreover, very little is known about the precise repartition
of ABPs at these scales in cells.
The effect of cooperativity at the single actin filament scale was progressively unveiled with the
development of real-time in vitro fluorescence microscopy. The effect of cooperativity is obvious when the
binding of ABPs is observed for various concentrations of ABPs. At a low concentration of ABP, bound
molecules are isolated. At an intermediate ABP concentration, cooperative effects are most important,
and ABPs begin to assemble on the side of actin filaments as clusters. The size and dynamics of these
clusters appears to vary widely depending on the nature of the ABPs. At a high concentration of ABP,
ABPs decorate actin filaments fully [13, 5]. The rapidity of these transitions depends on the amplitude
of cooperativity, which can be quantified by a Hill coefficient [18]. However, it is not determined to this
day how the amplitude and the range of cooperativity of an ABP affect the repartition of ABPs along
actin filaments at the molecular level.
For this study, our objective was to overcome current experimental limitations, by developing a
stochastic model of ABP cluster formation. This model allowed us to predict the behavior of ABP
cluster formation along actin filaments as function of the amplitude and the range of cooperativity.
We found that our model recapitulates the experimental behavior of proteins such as ADF/cofilin, and
allowed us to explore the consequences of cooperativity over a large range of parameters. Our results
indicate that the amplitude and the range of cooperativity modulate extensively actin binding protein
cluster size, density and dynamics. These findings suggest cooperativity as a key mechanism for cells to
control the repartition of ABPs to specific actin filament populations.
2
Materials and Methods
ADF/cofilin cluster formation and imaging in vitro
For the imaging of Alexa488-ADF/cofilin binding to Alexa568-labeled actin filaments, proteins were
purified, labeled and experiments performed as detailed in [13]. Experiments were performed in the
presence of 1 µM of globular actin and 3 µM of profilin and at concentrations of ADF/cofilin for which
severing events were the least frequent. Data were acquired on a Nikon Eclipse Ti microscope, equipped
with a 60X NA 1,49 objective, an OptoSplit II beam splitter and a Prime 95B scientific CMOS camera
(Photometrics).
Images were recorded using Metamorph software, and analyzed quantitatively with
ImageJ 1.49v. Average distances between clusters were evaluated by measuring the numbers of visible
clusters per total length of filamentous actin in a field of view 20 minutes after the initiation of the
experiment. This analysis was performed with the plugin Ridge Detection available for ImageJ. Values
indicated in the text are averages from 3 independent experiments.
Description of the model
To study clusters formation for a protein binding cooperatively to filamentous actin, we developed a
model that we simulate with a dynamic Monte-Carlo method. The studied system corresponds to
a single actin filament divided into equal compartments corresponding to sites where the ABPs can
bind. If not specified, our simulations are done with an actin filament of 8.104 subunits in a volume of
site, where lsite is the typical length of one site. We fixed lsite = 10 nm which corresponds to the
2.25.107l3
approximate size of an actin subunit. These values correspond to a concentration of [actin] (cid:39) 5.9 µM.
It models the case where the binding stoichiometry of the ABP to actin is 1:1, although results from
our work could easily be interpreted by extension to other binding stoichiometries (e.g. tropomyosin
which binds over the length of several actin subunits), or to any other 1 dimensional biological polymer
systems interacting with side binding ligands.
The algorithm proceeds as following. Each ABP can be in two different states: attached to the actin
filament or free in solution. When an ABP is bound to the filament, an energy is assigned stochastically
following a Boltzmann distribution, and the ABP unbinds when its energy exceeds the activation energy
of the corresponding chemical reaction. When an ABP is free in solution, the algorithm tests if the ABP
is in contact with the filament. The probability of contact with the filament is equal to the ratio between
the volume in which the ABP is in contact with actin and the total volume accessible to the ABP. When
an ABP is in contact with the actin filament, a compartment is chosen randomly. If the compartment is
already filled with another ABP, the ABP is left free in solution. On the contrary, if the compartment
is empty, an energy is assigned to the ABP similarly than before and compared to the activation energy
of binding. If this energy is larger than the activation energy, then the ABP binds.
We aspired to develop a model which would describe faithfully the biochemical properties of ABPs.
The main parameters accessible experimentally are reaction rates k, which are related to activation
energies Ea by the Arrhenius law
(1)
where A is an unknown prefactor, kB is the Boltzmann constant and T is the temperature. We ran
a set of preliminary simulations with different values of Ea. The rate of ABP binding to the actin
filament obtained from these simulations enabled us to compute the prefactor A, and to verify that
system follows the Arrhenius law (data not shown). These preliminary experiments enabled us to fix
the prefactor A for the subsequent simulations.
k = A exp(−Ea/kBT )
+
In our simulations, we fixed dissociation rates k− to a constant value along the actin filament, as
dissociation rates of most ABPs is not reported to be variable when ABPs dissociate spontaneously from
actin filaments and was even measured to be constant for cofilin [15]. On the contrary, binding rates of
take cooperativity into account in order to characterize how cooperativity may influence the
ABPs kcoop
dynamics and steady state regime of ABP cluster formation (Fig. 1-a)). Our model enables to study
the case of end-to-end cooperativity, but also allows to explore the possibility that allosteric interactions
along the actin filament change the spatial range of cooperativity beyond one actin subunit along the
in our model depends on the local distribution of
filament [3].
previously bound ABPs. The influence of cooperativity is defined by two parameters: ncoop, which
quantifies the typical number of sites over which the presence of a bound ABP has an influence on
binding rates; and Acoop, which is the amplitude of cooperativity (Fig. 1-b)). How cooperativity is
In other words, binding rates kcoop
+
3
i) on a
Figure 1: Illustrative figure explaining the model. a) Value of binding rate on the filament :
+, ii) on a decorated filament,
bare filament, the binding rate kcoop
of an ABP is influenced by the presence of other bound ABPs. In both cases the
the binding rate kcoop
dissociation rate is constant and equal to k−. b) The function f[ncoop,Acoop] defines the amplitude and
range of cooperativity at the site i when a single ABP is bound at the site n. Plots give examples of i)
short range and ii) long range cooperativities (see Eq.(3)).
of an ABP is constant and equal to kb
+
+
reinforced with the simultaneous binding of multiple ABPs has not been investigated experimentally to
our knowledge. Therefore, we chose a model where the contribution of all surrounding bound ABPs are
(see Fig.1-a)) and not only the nearest neighbor as previously studied [40, 6, 8, 5].
As a consequence, we consider the following binding rate
kcoop
+ (n) = kb
+ +
s(i)f[ncoop,Acoop](i − n)
(cid:88)
i
where
f[ncoop,Acoop](i − n) = Acoop
= Acoop
j(cid:54)=n exp(cid:2)−{(j − n)2 − 1}(cid:3)
(cid:80)
j(cid:54)=n exp(cid:2)−{(j − n)2 − 1}/n2
(cid:80)
(cid:3)
exp(cid:2)−{(i − n)2 − 1}/n2
coop
coop
.
Asum(ncoop)
(cid:3) exp(cid:2)−{(i − n)2 − 1}/n2
coop
(cid:3)
s(i) = 1 if the compartment is occupied and s(i) = 0 if not, n is the site number, Acoop is the amplitude
is defined to be, for a
of cooperativity and kb
given Acoop, the same binding rate at the end of an infinite cluster of ABPs independently of ncoop. Most
sections of the manuscript refer to the ratio Acoop/kb
+ which normalizes the amplitude of cooperativity
Acoop by the association constant kb
+.
+ is the binding rate of a bare actin filament (Fig. 1). kcoop
+
Results and Discussion
Validation with Hill Formalism
We first aimed at validating our approach by comparing the results from our model with a theoretical
description based on the Hill formalism. In this formalism the density of decorated actin filaments is
plotted with respect to the concentration of free ABPs. Binding curves display characteristic sigmoidal
shapes which represent a transition from barely decorated to fully decorated actin filaments. In the
frame of the Hill model [18], the steady state density of decorated actin filaments d typically follows the
function
(2)
(3)
(4)
d =
[free ABP]n
K ef f
d + [free ABP]n
4
where [free ABP] is the concentration of free ABPs, n is the Hill coefficient and K ef f
dissociation constant. The Eq. (4) can also be reformulated as
d
is an effective
d
1 − d
=
[free ABP]n
K ef f
d
(5)
which gives a power law and is more convenient to characterize. The Hill coefficient n quantifies the
degree of cooperativity and how abrupt is the transition while K ef f
We fixed for these simulations the value of Kd = 2.2.10−5 M, which is in the range of experimental
values for various ABPs binding cooperatively to actin (e.g, for cofilin, Kd (cid:39) 10−6 − 10−5 M [2, 6, 8, 15]
and for tropomyosin, Kd (cid:39) 10−6 − 10−3 M [41, 12, 5]). We performed simulations for two values of
+, 1 and 104, corresponding respectively to weak and strong cooperativities (Fig. 2). For each
Acoop/kb
value of Acoop, six different values of ncoop (1, 4, 6, 25, 50 and 100) were computed. Each data point
corresponds to the steady state values of d and d/(1− d) as a function of the concentration of free ABP
in solution.
locates this transition.
d
Figure 2: Evolution of the steady state density of decorated actin filaments d as a function of the
concentration of free ABP for different values of ncoop. The symbols represent simulations results.
Insets : evolution of d/(1 − d) in log-log scale. a) in the case of weak cooperativity (Acoop/kb
+ = 1),
the black solid line is the fit of Hill model for ncoop = 1, the pink solid line is the fit of Hill model
for ncoop = 100 and the dashed line is the corresponding curve if there is no cooperativity. b) In the
+ = 104), the solid lines represent the different fits of Hill model.
case of strong cooperativity (Acoop/kb
Parameters corresponding to fits of Hill model are shown in Table 1.
Results obtained from the simulations confirm that the model follows closely the theory described by
the Hill equation, although it diverges progressively at high concentration of ABPs for strong and long
range of cooperativities (Fig. 2). We can determine the Hill coefficients and the effective dissociation
constants by fitting the results from our simulations with Eqs (4) and (5) (see Table 1). In the weak
+ = 1), the Hill coefficient n is increased of 12 − 14% with respect to
cooperativity regime (Acoop/kb
a non-cooperative situation where n = 1. The effective dissociation constant is about 6.8 − 7.9 times
lower that the Kd on bare actin filaments, which means an enhancement of adhesion to the filament.
We note that in the case of weak cooperativity, it is rather difficult to extract a dependence with the
range of cooperativity ncoop. In the strong cooperativity regime (Acoop/kb
+ = 104), the effect is more
important. The Hill coefficient is multiplied by more than ten times. The value of n is now varying with
the range of cooperativity ncoop. It starts from about 14 at very short range (ncoop = 1) to reach about
18 for long range cooperativity. The effect on the effective dissociation constant K ef f
is more apparent.
is decreased from 113 order of magnitude at short range to about 160 order of magnitude for long
K ef f
range cooperativity. The evolution seems monotonous with the range of cooperativity ncoop, although
the results for ncoop = 50 deviates from the trend in this set of simulations.
Overall, the behavior of the model is consistent with the Hill model, and the calculated values of Hill
coefficients are consistent with the orders of magnitude that are published in the literature. The typical
measured values of n are in the range of 1 − 10 for ADF/cofilin [26, 6] and tropomyosin [16]. Another
study [28] suggests that n could be even much larger than 10 in case of tropomyosin.
d
d
5
Table 1: Hill model's parameters obtained by fitting simulations results shown in Fig.2.
(M )
(M ) Acoop/kb
+ = 1 ncoop
+ = 104
Acoop/kb
ncoop
1
4
6
25
50
100
n
13.922
15.459
18.099
18.611
15.667
18.124
d
K ef f
3.673.10−118
9.6237.10−135
1.366.10−158
7.668.10−165
3.7987.10−139
2.3431.10−161
n
1.1229
1.1309
1.138
1.1379
1.1285
1.131
d
K ef f
3.227.10−6
2.829.10−6
2.587.10−6
2.587.10−6
2.866.10−6
2.784.10−6
1
4
6
25
50
100
Impact of Cooperativity on the Length of ABP Clusters
The use of the model to understand how clusters of ABPs are formed required first a precise definition
of clusters. Defining clusters as strictly consecutive bound ABPs along an actin filament was not very
informative, as many clusters presented empty gaps due to the natural stochasticity of the process. This
strict definition also did not permit us to compare the results of the simulations with experimental results
obtained by fluorescence microscopy. In such assays, the limit of detection between 2 close molecules
along an actin filament is limited by diffraction. The minimal distance to distinguish 2 fluorescent
molecules emitting simultaneously is determined by the Rayleigh criterion, which corresponds to a
distance of about 200 nm for usual optical wavelengths. Better resolutions can be achieved with super-
resolution methods which typically decrease the limit of separation to distances of few tens of nanometers,
corresponding to an approximate distance lopt = 20 actin subunits ([?] and our unpublished data). We
used this reasonable criterion as a standard length to separate individual clusters in our model (Fig.
3-a).
A large range of concentration of ABPs was tested (see Fig. S1). Similar to what is observed in
vitro by fluorescence microscopy [13], low concentrations of ABPs triggered the formation of small and
localized clusters of ABPs along the actin filament, while larger concentrations induced a quite abrupt
transition to a fully decorated state. We then aimed at determining the influence of the amplitude of
c at steady state labeled by the index ∞ (see Fig.
cooperativity Acoop for the average cluster length l∞
3-b). The number of ABPs was fixed to 5000, which corresponds to a concentration of about 370 nM.
We started studying the effect of changing the affinity Kd of the ABP for a fixed range of cooperativity
ncoop = 1 (Fig. 3-b)). Two regimes can be identified. For an amplitude of cooperativity Acoop/kb
+ lower
than about 300, the affinity Kd has a dramatic effect on the average cluster length l∞
c . We identify this
regime as a weak cooperativity regime because in the limit of very small Acoop, the behavior is dominated
by non-cooperative binding. In this extreme case, the ABPs bind to the filament randomly in space
with approximately the same rates they would have on a bare filament. Moreover, the number of bound
ABPs increases for smaller dissociation constants, increasing the probability to form long clusters. In
the weak cooperativity regime, the length of the clusters lc is therefore mainly ruled by the affinity of
+ higher than about 300, the behavior becomes dominated by cooperative
the ABP Kd. For Acoop/kb
binding. We identify this regime as a strong cooperativity regime because the average cluster length
l∞
c at steady state does not depend on Kd. The total amount of bound ABPs and their distribution
along the filament is also independent of Kd (data not shown). The cluster length lc is not led by
probability, but rather by the amplitude of cooperativity Acoop. The results are similar for longer range
cooperativity: two regimes are observed with a boundary at the nearly same values (data not shown).
We can relate from Eq.(2) the probability to bind the adjacent sites of a bound ABP with the
characteristics of cooperativity for a given ABP corresponding to the boundary. Acoop/kb
+ = 300 means
that if ncoop = 1, an ABP will bind next to a single bound ABP 301 times faster than to a bare
filament.
If ncoop = 100, it will bind next to a single bound ABP about 4.6 times faster than to a
bare filament. In both situations, an ABP will bind next to a long decorated portion 316 times faster
compared to a bare filament. For proteins like tropomyosin, where ncoop is believed to be equal to 1,
values reported in literature indicate an increase of the binding probability of (cid:39) 102 − 103 times to the
sites that are adjacent to an already bound tropomyosin compared to a random site along the filament
[41, 19, 40]. These values are sufficient to characterize the parameters of cooperativity, and suggest
that tropomyosin is at the boundary between the weak and strong cooperativity regimes. However,
for ADF/cofilin, reported values indicate a 2.3 times faster binding rate in the vicinity of an already
bound cofilin [15], but the uncertainty in the range of cooperativity ncoop does not permit to evaluate
6
c , l∞
c and distance between clusters l∞
Figure 3: Evolution of the steady state length of ABP clusters l∞
for different cooperative behaviors. Values of l∞
empty and ncoop are given in number of adhesion sites.
a) Schematic cartoon defining the different length notations used in this study. The ∞ symbol indicates
steady state values. b) Average cluster length l∞
c as a fonction of the amplitude of cooperativity Acoop/kb
+
for different values of affinity Kd and for a range of cooperativity ncoop = 1. The Kd represented by open
circles is equal to 2.2.10−5 M (value used in c), d) and e)). The vertical dashed line delimits the weak
cooperativity regime from the strong cooperativity regime. c) Average cluster length l∞
c as a fonction of
+ for different ranges of cooperativity ncoop. d) Average distance
the amplitude of cooperativity Acoop/kb
between clusters l∞
empty for different ranges of cooperativity ncoop e) Ratio between the average distance
between clusters l∞
empty and the average cluster length l∞
c as a fonction of the amplitude of cooperativity
+.
Acoop/kb
empty
unambiguously the amplitude of cooperativity Acoop.
We then aimed at determining the influence of the range of cooperativity for the average cluster
c . We fixed in these simulations a value of Kd = 2.2.10−5 M and evaluated the
length at steady state l∞
effect of modulating the range of cooperativity ncoop (Fig. 3-c)). The behavior for small Acoop is mainly
ruled by the affinity Kd and not by cooperativity, and explains the formation of very short clusters. We
will therefore mainly focus on the strong cooperativity regime. The first observation is that the cluster
length l∞
increases
increases with the amplitude of cooperativity Acoop. For 102 < Acoop/kb
+ < 105, l∞
c
c
7
with the range of cooperativity ncoop until ncoop = 25 binding sites, but decreases for values of ncoop
greater than 25 binding sites. This means that there is an optimum range of cooperativity to form long
clusters. In the limit of short range cooperativity, ABPs will be progressively recruited to the nearest
sites available and form short clusters.
In the opposite limit of long range cooperativity, ABPs will
have the opportunity to bind further away along the actin filament, which will increase the likelihood to
form new independent clusters instead of growing the original cluster. A second observation is that l∞
c
seems to reach a plateau in the limit of very strong cooperativity, and the asymptotic values decrease
with increasing values of ncoop. Some additional points would be necessary for large values of Acoop
to confirm rigorously this assertion, but very large values of Acoop are probably unrealistic and would
require very long simulation times.
In addition to cluster length, we noticed that cooperativity also modulates the distribution of clusters
along actin filaments. To study this effect, we calculated l∞
empty, which is the average distance between
+ (Fig. 3-d)). As for cluster length
two consecutive clusters at steady state, as a function of Acoop/kb
l∞
c , l∞
empty does not change with ncoop for small Acoop where ABP binding is mainly ruled by the affinity
+ < 105, there is an optimal range of cooperativity
Kd and not by cooperativity. For 102 < Acoop/kb
ncoop = 4 − 6 for which clusters are most distant from one another. Moreover, l∞
empty increases in
the strong cooperativity regime until it probably reaches a plateau under the limit of very strong
cooperativity. This result indicates that the formation of long clusters is correlated with the maintenance
of long sections of bare filaments between clusters. In this regime, it is more probable to form long clusters
than to create new ones. The ratio between the distance between consecutive clusters and their length
remains rather constant in the strong cooperativity regime (Fig.3-e)). This asymptotic value decreases
with the range of cooperativity ncoop from about 6-7 for ncoop = 1 to about 2 for ncoop = 100. In the
case of a long range cooperativity, l∞
empty are both quite small and of similar values (about 35 and
55 adhesion sites respectively). This result indicates that a long range cooperativity is unable to form
isolated clusters with the current definition, and cannot account for observations made by fluorescence
microscopy [13].
c and l∞
To summarize this section, our model predicts that for cooperativity amplitudes 102 < Acoop/kb
+ <
105, different optimum ranges of cooperativity ncoop are predicted to obtain specific characteristics of
clusters. A range of cooperativity of ncoop = 25 is optimal to form long clusters of ABPs, while a range
of cooperativity of ncoop = 4 − 6 is optimal to form distant clusters of ABPs. However, it is for the
shortest ranges of cooperativity (ncoop = 1) that the distance between cluster is the longest with respect
to their length. On the contrary, very long range cooperativities are ineffective to form isolated clusters.
ABP
Impact of Cooperativity on the Number of ABPs per Cluster
As we previously defined a length lopt = 20 empty adhesion sites above which clusters are optically
separated, the length of the clusters is not correlated with the number of ABPs per cluster. In this
section, we focus on the impact of cooperativity in modulating the ABP density within individual
clusters at steady state.
(Fig. 4-a)). In the strong cooperativity regime, (N cluster
We first investigated how the amplitude of cooperativity Acoop/kb
+ influences the number of ABPs per
ABP )∞ increases with Acoop/kb
cluster N cluster
+
until it reaches an asymptotic value which decreases with ncoop (as mentioned previously, additional
simulations would be necessary to confirm formally the existence of the plateaus). The typical maximum
values are below 30 bound ABPs, which is much smaller than total number of 5000 ABPs; therefore,
plateaus are not due to finite-size effects. For 102 < Acoop/kb
+ < 103, the number of ABPs per cluster
is maximum for a range of cooperativity ncoop between 6 and 25. The non monotonous evolution of
ABP )∞ as a function of ncoop indicates an optimal range of cooperativity to accumulate proteins
(N cluster
ABP )∞ saturates faster for long ranges of cooperativity. As
within clusters. For larger Acoop/kb
+, (N cluster
a consequence, for 103 < Acoop/kb
+ < 106, the optimum range of cooperativity ncoop is lower, between
4 and 6.
In contrast, the curve of ncoop = 1 crosses over all the other curves. This indicates that
end-to-end cooperativity is an efficient way to recruit a large number of ABPs per clusters, as long as
the amplitude of cooperativity is strong enough (i.e. 106 < Acoop/kb
+) to drive the process. Conversely,
a long range cooperativity is ineffective to recruit many ABPs within one cluster at any Acoop. For
example, the maximum (N cluster
ABP )∞ is 5.6 for ncoop = 100.
We can derive the density of bound ABPs per cluster at steady state defined as the ratio between
(Fig. 4-b)). Strikingly, these
the number of ABPs per cluster (N cluster
ABP )∞ and the cluster length l∞
c
8
ABP )∞ as a fonction of the
Figure 4: a) Evolution of the steady state number of ABPs per cluster (N cluster
+ for different ranges of cooperativity ncoop. b) Density of ABPs per
amplitude of cooperativity Acoop/kb
cluster (N cluster
+ for different ranges of
cooperativity ncoop. The dashed lines indicate the boundary between weak and strong cooperativity
regimes.
c as a fonction of the amplitude of cooperativity Acoop/kb
ABP )∞/l∞
curves evolve in opposite directions when approaching the strong cooperativity regime. While ABP
+ for long range cooperativity (i.e. ncoop = 25, 50 or 100),
density per cluster decreases with Acoop/kb
ABP density increases for short range cooperativity. The reason is that for large ranges of cooperativity
ncoop, ABPs can be stochastically recruited at larger distances than the average distance between two
consecutive bound ABPs within one cluster. As a consequence, the density declines. By contrast, for
small ncoop, ABPs bind at distances which is shorter than the typical distance between ABPs within
clusters, therefore the ABP density grows. Moreover, a plateau is reached at larger Acoop/kb
+ for all
ranges of cooperativity. The values of this plateau in the limit of very strong cooperativities decrease
with ncoop, indicating that a high density of protein per cluster requires necessarily a short range of
cooperativity.
+ < 106, a
range of cooperativity of ncoop = 4 − 6 is optimal to recruit many ABPs per cluster. Only for stronger
cooperativities is the limit of very short range of cooperativity ncoop = 1 better. Regarding the ABP
density of these clusters, the limit of very short range cooperativity is systematically optimal. On the
contrary, long range cooperativities are inefficient to recruit many ABPs per clusters and to form dense
clusters.
To summarize this section, our model predicts that for cooperativity amplitudes Acoop/kb
Impact of Cooperativity on Clusters Rate of Assembly
While results presented in the previous sections correspond to cluster configurations at steady state, this
section focuses on their dynamics of assembly. A dimensionless time ¯t is expressed in number of 1/k−
and all ABPs are unbound at ¯t = 0. In the case of ADF/cofilin, k− is reported to be 5.10−3− 5.10−1 s−1
[2, 43], so the unit for ¯t is 2 − 200 s. For tropomyosin, the dynamics is faster as k− = 102 − 103 s−1
[42, 40, 5], leading to time unit of 10−3 − 10−2 s. Nevertheless, the end-to-end cooperativity may
decrease the unbinding rate of one or two orders of magnitude for tropomyosin. Additionally, to avoid
the noise of early dynamics due to finite size effects, we simulated a larger system of 40000 ABPs for
640000 adhesion sites, while keeping the same concentrations and dissociation constants as in previous
sections.
We analyzed the growth and dynamics of ABPs clusters from the binding of the first molecules until
steady state. The total amount of bound ABPs increases smoothly until equilibrium amount is attained
(data not shown). In the meantime, both average length of clusters lc and average number of ABPs
also increase with similar timescales. Interestingly, a maximum value appears for
per cluster N cluster
sufficiently large Acoop/kb
+. This means they then decrease slightly toward their steady state values (see
Fig. 5-a)). The reason is that at early time and for large amplitude of cooperativity, the binding in
vicinity of existing clusters is much faster than unbinding. As a consequence, almost all binding events
ABP
9
ABP
over time. b) Average time ¯tmax
ABP )max as a function of the amplitude of cooperativity Acoop/kb
Figure 5: Dynamics of cluster assembly. Time is expressed in number of 1/k− and cluster length lc
in number of binding sites. a) Evolution of average cluster length lc and average number of ABP per
cluster N cluster
for clusters to reach the highest number of bound
proteins (N cluster
+. c) Average time ¯tmax
as a function of the amplitude of cooperativity Acoop/kb
for clusters to reach the longest size lmax
+.
ABP )max as a function of the amplitude of cooperativity.
d) Maximum number of bound proteins (N cluster
e) Maximum length of the clusters (N cluster
ABP )max as a function of the amplitude of cooperativity. In
b,c,d,e), the dotted black lines represent power functions with exponents corresponding to the fits of the
surrounding curves.
N
c
l
ABP
contribute to amplify N cluster
(and lc) as long as available reservoir of ABPs is large enough. When this
reservoir becomes limited, meaning that equilibrium amount of bound ABPs is attained, the number
of unbinding events is no longer negligible and the dynamics of clusters is ruled by the substitution of
bound ABPs rather than the addition of new ones. The likelihood that individual clusters split into two
different clusters of smaller size is more critical and implies decrease of both N cluster
and lc.
Interestingly, the time ¯tmax
to reach the maximum number of proteins per cluster (N cluster
ABP
N
ABP )max,
10
l
l
c
when existing, is systematically shorter than the time ¯tmax
to reach the maximum length of clusters lmax
(see Fig. 5-a,b,c)). The reason is that at ¯tmax
N , clusters are dense, particularly for small values of ncoop
(see Fig. S2), and only few adhesion sites are available except on the side of the clusters. Meanwhile,
unbinding events continue to occur within clusters, diminishing (N cluster
ABP ). Moreover, as a consequence
of cooperativity, the binding rate in the vicinity of clusters is large. Therefore, additional ABPs will
bind preferentially outside clusters and extend them, which means that the average length of clusters
lc continues to increase. As a consequence, ¯tmax
N . It is also of interest to mention that typical
distances separating clusters lempty are larger than every ncoop used in this study for large Acoop/kb
+
(see Fig. S2). This means that probability to merge two successive clusters is quite small.
≥ ¯tmax
We analyzed then how the parameters of cooperativity impact rates of cluster assembly. We first
N decreases
observed that the average time needed for clusters to reach their maximal number of ABP ¯tmax
with the amplitude of cooperativity Acoop/kb
+ following power laws (see Fig. 5-b)). The reduction is
explained by the fact that although clusters reach a higher number of bound ABP when Acoop/kb
+
increases, the binding rate of ABPs increases faster with the amplitude of cooperativity. Furthermore,
the time ¯tmax
diminishes with the range of cooperativity ncoop. The reason is that for short ranges of
cooperativity, recruitment rates of new ABPs are slower because limited to lower numbers of binding
ABP )max increases for ncoop ≤ 25
sites. Nevertheless, the maximum number of protein per cluster (N cluster
(see Fig. 5-d)). The effect of ncoop is therefore non intuitive as smaller clusters could be expected to
assemble faster. This is not the case, and we can conclude from these results that increasing the range of
cooperativity up to 25 subunits along an actin filament is a good strategy to have simultaneously a fast
clustering and a large number of bound ABPs. In contrast, both the maximum number of bound ABP
ABP )max and the average time needed to reach this threashold decline for ncoop ≥ 25. Overall, the
(N cluster
optimal value to form populated clusters at steady state is ncoop = 4 − 6 while it is ncoop = 25 at ¯tmax
(see Fig. 4-a) and Fig. 5-d)), although clusters are less dense for ncoop = 25 than for ncoop = 4 − 6 at
short time, when unbinding is negligible (see Fig. S2).
N
N
l
N . However, when Acoop/kb
+ in regime where maximums exist, the average
For the smallest amplitudes of cooperativity Acoop/kb
follows the same trend than the average time
time required to reach the maximal length of clusters ¯tmax
+ increases,
required to reach the maximal number of ABP per cluster ¯tmax
progressively diverges to eventually increase and follow a power law (see Fig. 5-b,c)). As previously
¯tmax
l
explained, when the equilibrium amount of bound ABPs is attained, the elongation of clusters at their
extremities is mostly due to the substitution than to the addition of new ABPs. Therefore, when
+ increases, the binding rate of ABP is more important in the vicinity of the clusters than in
Acoop/kb
bare sections of the filament, and new binding events enlarge existing clusters. This is consistent with
+ (see Fig. S2). Restriction of new binding area slows
an increase of cluster density at ¯tmax
with Acoop/kb
. This effect is even stronger for short range cooperativities where
down elongation and increases ¯tmax
this restriction is more important. As a result, ¯tmax
+ for strong cooperativity and
decreases with the range of cooperativity ncoop. Finally, the maximum cluster length lmax
grows with
+ following a power law with an exponent which is independent
the amplitude of cooperativity Acoop/kb
(Fig. 5-d),
of the range of cooperativity ncoop (Fig. 5-e). This exponent is smaller than for N cluster
leading to denser clusters in the strong cooperativity limit. Moreover, similar to the steady state value
l∞
c , the maximum
c , the maximum cluster length lmax
value lmax
does not seem to saturate in limit of very strong cooperativity. This absence of saturation is
also observed for N cluster
ABP .
is optimal for ncoop = 25. Though, inverse to l∞
increases with Acoop/kb
To summarize this part, the assembly rate of the ABP clusters is correlated with the range of
cooperativity ncoop. The shorter is the range, the slower is the dynamics. Then, a maximum value
+ is large enough. The time required to reach the maximum
appears for lc and N cluster
number of ABPs per clusters decreases with the amplitude of cooperativity Acoop/kb
+, while the time to
reach the maximum cluster length increases for strong cooperativity.
ABP when Acoop/kb
ABP
c
c
c
l
l
l
System Configurations at Fixed Number of ABPs per Clusters
We now focus on a well-described family of proteins, ADF/cofilins, as an example of how the model can be
used to predict the characteristics of cooperativity of an ABP. The cooperativity of ADF/cofilin binding
to actin filaments has been reported extensively, but the measured amplitude and range of cooperativity
along actin filaments varies considerably between studies. Microscopy experiments previously revealed
the behavior of ADF/cofilin binding to individual actin filaments. Above a minimal concentration
of ADF/cofilin in solution, the protein binds visibly to actin filaments and forms stable clusters of
11
reproductible size at the 10 seconds timescale [13]. We measured experimentally for this study the
average distance between clusters at three different concentrations of ADF/cofilin (Fig. 6-a,b) and Fig.
S3). Measured values vary from 6 to 42 µm, which corresponds to 2000 to 14000 subunits of actin. The
order of magnitude between experimental measurements and the model are comparable. Moreover, the
trend is similar, as the simulations predict that lempty decreases with the concentration of ABPs (see
Fig. S1-c,d)).
Then the number of bound molecules was evaluated by quantitative fluorescence microscopy, and
was found to remain below 10 molecules per cluster for a concentration of 180 nM of ADF/cofilin, and
to reach on average 23 molecules per cluster for a concentration of 360 nM of ADF/cofilin. When the
threshold of 23 molecules is reached, these isolated clusters are triggering the polarized severing of the
actin filaments, towards their pointed ends, at the interface between bare and decorated sections of
the filament. At higher concentration of ADF/cofilin, the protein fully decorates, copolymerizes with
actin and stabilizes the filaments [13]. While previous simulations were run with a constant pool of
ABPs, we varied in the new simulations the pool of ABPs. We determined the cooperativity conditions
leading to the formations of experimentally observed clusters of 10 or 23 bound proteins on average,
and analyzed for these configurations the characteristics of the clusters (Fig. 6). We generated results
for 2 representative amplitudes of cooperativity Acoop/kb
+ = 1 and 104, which corresponded respectively
to the weak and strong cooperativity regimes.
A first general observation is that all values are weakly influenced by the range of cooperativity
ncoop in the weak cooperativity regime. The length of clusters in this regime is always longer than
the distance lempty separating them. This means that clustering observed in experiments with isolated
clusters of ADF/cofilin is not compatible with the weak cooperativity regime. In contrast, in strong
cooperativity regime, the range of cooperativity ncoop has a strong influence in case of 10 ABPs per
cluster. The average distance between clusters increases from about 300 adhesion sites between clusters
for ncoop = 1 to about 1000 adhesion sites (the order of magnitude measured experimentally) for the
optimum value ncoop = 6. Above ncoop = 6, it decreases progressively to reach the same level than the
weak cooperativity regime for the longest ranges of cooperativity (Fig. 6-d). Evolution for clusters of 23
ABPs is similar with smaller lempty (about 100 adhesion sites for maximum value). These results exclude
in the case of ADF/cofilin the possibility of a range of cooperativity much above 25 actin subunits, and
makes a very short range of cooperativity much less likely than values around ncoop = 4 − 6.
Another important experimental observation is that the size of ADF/cofilin clusters remains rather
small [13]. They systematically appear as spots by fluorescence microscopy, which strongly suggests
that all molecules are not distinguishable optically and located within a section of less than about 200
nm, which corresponds to about 20 adhesion sites. In the weak cooperativity regime, the average cluster
length l∞
is systematically greater than 70 adhesion sites (Fig. 6-e). In the strong cooperativity regime,
our simulations indicate on the contrary that below a range of cooperativity ncoop = 10, clusters of ABPs
are short enough to account for the experimental data in case of 10 ABPs per cluster. Nevertheless, for
cluster of 23 ABPs, simulations values seem a little bit overestimated. This result suggests again that
the larger ranges of cooperativity are less probable in the case of ADF/cofilin.
We also investigated the concentration of ABPs which is necessary to form these clusters (Fig.
6-f). As mentioned above, in weak cooperativity regime, the formation of clusters is independent
of the concentration of ABPs. The typical concentrations required to form clusters in this regime
are 2.8 and 3.8 µM respectively for small and large clusters (or about half the actin concentration),
which is well above what is needed in vitro. The required amount of protein is smaller in the strong
cooperativity regime, and the effect is non monotonous with the range of cooperativity ncoop. The
smallest required amount of ADF/cofilin is predicted for a range of cooperativity of ncoop = 6. For this
range of cooperativity, clusters of 23 ABPs are formed at a concentration of [ADF/cof ilin] (cid:39) 0.7 µM
((cid:39) 0.1[actin]), and clusters of 10 ABPs are formed at a concentration of [ADF/cof ilin] (cid:39) 50 nM
((cid:39) 0.008[actin]). These concentrations are in relatively good agreement with concentrations used in
the experiments. Typical time to reach maximum number of ABPs per cluster at Acoop/kb
+ = 104 and
ncoop = 6 is 0.2/k−. For cofilin, k− = 5.10−3 − 5.10−1 s−1 [2, 43], so typical timescales are 0.4 − 40 s.
This is in agreement with the time needed to reach clusters of 23 cofilin in experiments [13].
c
To summarize this part, evolution of lempty in simulations is consistent with experiments even if
strict comparison is difficult here. Then, the model suggest that in case of ADF/cofilin, long range
and/or weak cooperativity unlikely.
12
empty, l∞
Figure 6: Predicted configurations of the system to generate clusters of 10 and 23 ABPs on average
at steady state. Values of l∞
c and ncoop are given in number of adhesion sites. Each datapoint
indicates the values obtained from the closest simulations. Curves link average values. a) Example of
multiple Alexa-488-labeled ADF/cofilin (67 nM; in green) clusters formed on the side of an individual
actin filament (in red). Scale bar=3 µm. b) Experimental values measured from a).
c) Cartoon
explaining the significance of symbols. Closed symbols (resp. opened) are results of simulations for the
weak (resp. strong) cooperativity regime. d) Average distance between successive clusters l∞
empty as a
function of the range of cooperativity ncoop. e) Average cluster length l∞
c as a function of the range of
cooperativity ncoop. f) Average concentration of ABP required to obtain clusters of 10 and 23 molecules
as a function of the range of cooperativity ncoop.
Conclusion
In this study, we have detailed the impact of binding cooperativity for the formation of protein clusters
to the side of individual actin filaments. We have principally investigated the influence of two parameters
which are the amplitude and the range of cooperativity. We found that both parameters have a strong
influence on cluster assembly properties such as size, density and dynamics of the clusters, and these
effects are summarized in Fig. 7. These properties are impacted differently by the parameters of
cooperativity. For example, building dense clusters of ABPs at a fast rate will require trade-off values of
13
the amplitude and range of cooperativity. As ABPs have been demonstrated experimentally to follow a
diversity of cooperative binding behaviors, it is expected that cooperativity is used in cells as a powerful
mechanism to control the decoration of actin filaments. Depending on the function of the ABP and the
level of activity required for a specific population of actin filaments, cells will favor locally the formation
of clusters of appropriate size, density and dynamics.
Figure 7: Cartoon summarizing how the amplitude and range of cooperativity impact ABP cluster
formation. Each arrow indicates how specific properties of cluster formation are enhanced from their
lower values to their higher values in this 2-dimensional diagram.
Author Contributions
T.L.G. and A.M. conceived the project and wrote the manuscript. T.L.G. designed the model, performed
the simulations, analyzed the computational results and compared with the experimental results. A.M.
performed and analyzed the experiments and funded the project.
Acknowledgments
The authors would like to thank Christopher P. Toret for his critical reading of the manuscript. This
project has received funding from the European Research Council (ERC) under the European Union's
Horizon 2020 research and innovation programme (grant agreement no 638376/Segregactin).
14
References
[1] L. Blanchoin, R. Boujemaa-Paterski, C. Sykes, and J. Plastino. Actin dynamics, architecture, and
mechanics in cell motility. Physiol. Rev., 94:235 -- 263, 2014.
[2] L. Blanchoin and T. D. Pollard. Mechanism of interaction of acanthamoeba actophorin (adf/cofilin)
with actin filaments. J. Biol. Chem., 274:15538 -- 15546, 1999.
[3] A. A. Bobkov, A. Muhlrad, D. A. Pavlov, K. Kokabi, A. Yilmaz, and E. Reisler. Cooperative effects
of cofilin (adf) on actin structure suggest allosteric mechanism of cofilin function. J. Mol. Biol.,
356:325 -- 334, 2006.
[4] W. Cao, J. P. Goodarzi, and E. M. De La Cruz. Energetics and kinetics of cooperative cofilin -- actin
filament interactions. J. Mol. Biol., 361:257 -- 267, 2006.
[5] J. R. Christensen, G. M. Hocky, K. E. Homa, A. N. Morganthaler, S. E. Hitchcock-DeGregori,
G. A. Voth, and D. R Kovar. Competition between tropomyosin, fimbrin, and adf/cofilin drives
their sorting to distinct actin filament networks. eLife, 6, 2017.
[6] E. M. De La Cruz. Cofilin binding to muscle and non-muscle actin filaments: isoform-dependent
cooperative interaction. J. Mol. Biol., 346:557 -- 564, 2005.
[7] E. M. De La Cruz. How cofilin severs an actin filament. Biophys. Rev., 1:51 -- 59, 2009.
[8] E. M. De La Cruz and D. Sept. The kinetics of cooperative cofilin binding reveals two states of the
cofilin-actin filament. Biophys. J., 98:1893 -- 1901, 2010.
[9] V. E. Galkin, A. Orlova, D. S. Kudryashov, A. Solodukhin, E. Reisler, G. F. Schröder, and E. H.
Egelman. Remodeling of actin filaments by adf/cofilin proteins. Proc. Natl. Acad. Sci. USA,
108:20568 -- 20572, 2011.
[10] V. E. Galkin, A. Orlova, N. Lukoyanova, W. Wriggers, and E. H. Egelman. Actin depolymerizing
factor stabilizes an existing state of f-actin and can change the tilt of f-actin subunits. J. Cell Biol.,
153:75 -- 86, 2001.
[11] V. E. Galkin, A. Orlova, G. F. Schröder, and E. H. Egelman. Structural polymorphism in f-actin.
Nat. Struct. Mol. Biol., 17:1318 -- 1324, 2010.
[12] G. Gateva, E. Kremneva, T. Reindl, T. Kotila, K. Kogan, L. Gressin, P. W. Gunning, D. J.
Manstein, A. Michelot, and P. Lappalainen. Tropomyosin isoforms specify functionally distinct
actin filament populations in vitro. Curr. Biol., 27:705 -- 713, 2017.
[13] L. Gressin, A. Guillotin, C. Guérin, L. Blanchoin, and A. Michelot. Architecture dependence of
actin filament network disassembly. Curr. Biol., 25(11):1437 -- 1447, 2015.
[14] P. Gunning, G. O'neill, and E. Hardeman. Tropomyosin-based regulation of the actin cytoskeleton
in time and space. Physiol. Rev., 88:1 -- 35, 2008.
[15] K. Hayakawa, S. Sakakibara, M. Sokabe, and H. Tatsumi. Single-molecule imaging and kinetic
analysis of cooperative cofilin-actin filament interactions. Proc. Natl. Acad. Sci. USA, 111:9810 --
9815, 2014.
[16] R. W. Heald and S. E. Hitchcock-DeGregori. The structure of the amino terminus of tropomyosin is
critical for binding to actin in the absence and presence of troponin. J. Biol. Chem., 263:5254 -- 5259,
1988.
[17] G. Hild, B. Bugyi, and M. Nyitrai. Conformational dynamics of actin: effectors and implications
for biological function. Cytoskeleton, 67:609 -- 629, 2010.
[18] A. V. Hill. The combinations of haemoglobin with oxygen and with carbon monoxide. i. Biochem.
J., 7:471 -- 480, 1913.
15
[19] L. E. Hill, J. P. Mehegan, C. A. Butters, and L. S. Tobacman. Analysis of troponin-tropomyosin
binding to actin. troponin does not promote interactions between tropomyosin molecules. J. Biol.
Chem., 267:16106 -- 16113, 1992.
[20] A. Huehn, W. Cao, W. A. Elam, X. Liu, E. M. De La Cruz, and C. V. Sindelar. The actin filament
twist changes abruptly at boundaries between bare and cofilin-decorated segments. J. Biol. Chem.,
293:5377 -- 5383, 2018.
[21] M. K. Jensen, E. J. Morris, R. Huang, G. Rebowski, R. Dominguez, D. A. Weitz, J. R. Moore, and
C.-L. A. Wang. The conformational state of actin filaments regulates branching by actin-related
protein 2/3 (arp2/3) complex. J. Biol. Chem., 287:31447 -- 31453, 2012.
[22] T. Kamasaki, R. Arai, M. Osumi, and I. Mabuchi. Directionality of f-actin cables changes during
the fission yeast cell cycle. Nat. Cell Biol., 7:916 -- 917, 2005.
[23] D. R. Kovar, V. Sirotkin, and M. Lord. Three's company: the fission yeast actin cytoskeleton.
Trends Cell Biol., 21:177 -- 187, 2010.
[24] J. Kozuka, H. Yokota, Y. Arai, Y. Ishii, and T. Yanagida. Dynamic polymorphism of single actin
molecules in the actin filament. Nat. Chem. Biol., 2:83 -- 86, 2006.
[25] B. R. McCullough, L. Blanchoin, J.-L. Martiel, and E. M. De La Cruz. Cofilin increases the
bending flexibility of actin filaments: implications for severing and cell mechanics. J. Mol. Biol.,
381:550 -- 558, 2008.
[26] A. McGough, B. Pope, W. Chiu, and A. Weeds. Cofilin changes the twist of f-actin: implications
for actin filament dynamics and cellular function. J. Cell Biol., 138:771 -- 781, 1997.
[27] A. Michelot and D. G. Drubin. Building distinct actin filament networks in a common cytoplasm.
Curr. Biol., 21:R560 -- R569, 2011.
[28] J. Moraczewska, K. Nicholson-Flynn, and S. E. Hitchcock-DeGregori. The ends of tropomyosin are
major determinants of actin affinity and myosin subfragment 1-induced binding to f-actin in the
open state. Biochemistry, 38:15885 -- 15892, 1999.
[29] K. X. Ngo, N. Kodera, E. Katayama, T. Ando, and T. Uyed. Cofilin-induced unidirectional coop-
erative conformational changes in actin filaments revealed by high-speed atomic force microscopy.
elife, 4:e04806, 2015.
[30] A. Orlova, E. Prochniewicz, and E. H. Egelman. Structural dynamics of f-actin: Ii. cooperativity
in structural transitions. J. Mol. Biol., 245:598 -- 607, 1995.
[31] T. D. Pollard. Actin and actin binding proteins. Cold Spring Harb. Perspect. Biol., 8:a018226,
2016.
[32] E. Prochniewicz, N. Janson, D. D. Thomas, and E. M. De La Cruz. Cofilin increases the torsional
flexibility and dynamics of actin filaments. J. Mol. Biol., 353:990 -- 1000, 2005.
[33] I. Rouiller, X.-P. Xu, K. J. Amann, C. Egile, S. Nickell, D. Nicastro, R. Li, T. D. Pollard, N. Volk-
mann, and D. Hanein. The structural basis of actin filament branching by the arp2/3 complex. J.
Cell Biol., 180:887 -- 895, 2008.
[34] V. Sirotkin, J. Berro, K. Macmillan, L. Zhao, and T. D. Pollard. Quantitative analysis of the
mechanism of endocytic actin patch assembly and disassembly in fission yeast. Mol. Biol. Cell,
21:2894 -- 2904, 2010.
[35] C. T. Skau and C. M. Waterman. Specification of architecture and function of actin structures by
actin nucleation factors. Ann. Rev. Biophys., 44:285 -- 310, 2015.
[36] T. M. Svitkina and G. G. Borisy. Arp2/3 complex and actin depolymerizing factor/cofilin in den-
dritic organization and treadmilling of actin filament array in lamellipodia. J. Cell Biol., 145:1009 --
1026, 1999.
16
[37] T. M. Svitkina, E. A. Bulanova, O. Y. Chaga, D. M. Vignjevic, S.-I. Kojima, J. M. Vasiliev, and
G. G. Borisy. Mechanism of filopodia initiation by reorganization of a dendritic network. J. Cell
Biol., 160:409 -- 421, 2003.
[38] K. Tanaka, S. Takeda, K. Mitsuoka, T. Oda, C. Kimura-Sakiyama, Y. Maéda, and A. Narita.
Structural basis for cofilin binding and actin filament disassembly. Nat. Commun., 9:1860, 2018.
[39] L. S. Tobacman. Cooperative binding of tropomyosin to actin. Adv. Exp. Med. Biol., 644:85 -- 94,
2008.
[40] A. Vilfan. The binding dynamics of tropomyosin on actin. Biophys. J., 81:3146 -- 3155, 2001.
[41] A. Wegner. Equilibrium of the actin-tropomyosin interaction. J. Mol. Biol., 131:339 -- 353, 1979.
[42] C. Weigt, A. Wegner, and M. Koch. Rate and mechanism of the assembly of tropomyosin with
actin filaments. Biochemistry, 30:10700 -- 10707, 1991.
[43] H. Wioland, B. Guichard, Y. Senju, S. Myram, P. Lappalainen, A. Jégou, and G. Romet-Lemonne.
Adf/cofilin accelerates actin dynamics by severing filaments and promoting their depolymerization
at both ends. Curr. Biol., 27:1956 -- 1967, 2017.
[44] M. E. Young, J. A. Cooper, and P. C. Bridgman. Yeast actin patches are networks of branched
actin filaments. J. Cell Biol., 166:629 -- 635, 2004.
17
Supplementary Material
Figure S1: Evolution of the steady state clusters characteristics (a) and b) average cluster length; c)
and d) average distance between clusters ; e) and f) ratio between the distance between clusters l∞
and the cluster length l∞
c ; g) and h) number of ABP per cluster) as a function of the concentration
of ABP. The filament length is 8.104 binding sites. Left plots correspond to the weak cooperativity
regime Acoop/kb
+ = 104.
Vertical dotted lines represent conditions equivalent to 5000 and 80000 ABPs respectively in a volume
of 2.25.107l3
site. The horizontal dotted lines in g) and h) correspond to clusters of 10 and 23 molecules.
+ = 1. Right plots correspond to the strong cooperativity regime Acoop/kb
empty
18
Figure S2: Evolution of clusters characteristics (cluster length, distance between clusters, ratio between
the distance between clusters and the cluster length, number of ABP per cluster and density of ABP
per cluster) as a function of the amplitude of cooperativity Acoop/kb
+. Left plots are data when system
reach the highest number of proteins per clusters (N cluster
ABP )max; center plots are data when clusters
reach the longest size lmax
; right plots are data at steady state.
c
19
Figure S3: Distribution of Alexa-488-labeled ADF/cofilin clusters (in green) along actin filaments (in
red) for three concentrations of ADF/cofilin 20 min after the initiation of the experiment. Scale bar
: 10 µm. Numbers of clusters per unit length of actin varies from 0.025 clusters per µm at 33 nM of
ADF/cofilin, to 0.076 clusters per µm at 50 nM of ADF/cofilin and to 0.149 clusters per µm at 67 nM
of ADF/cofilin.
20
33 nM ADF/cofilin50 nM ADF/cofilin67 nM ADF/cofilin |
1707.07832 | 3 | 1707 | 2018-06-15T13:47:34 | Transcription factor target site search and gene regulation in a background of unspecific binding sites | [
"physics.bio-ph"
] | Response time and transcription level are vital parameters of gene regulation. They depend on how fast transcription factors (TFs) find and how efficient they occupy their specific target sites. It is well known that target site search is accelerated by TF binding to and sliding along unspecific DNA and that unspecific associations alter the occupation frequency of a gene. However, whether target site search time and occupation frequency can be optimized simultaneously is mostly unclear. We developed a transparent and intuitively accessible state-based formalism to calculate search times to target sites on and occupation frequencies of promoters of arbitrary state structure. Our formalism is based on dissociation rate constants experimentally accessible in live cell experiments. To demonstrate our approach, we consider promoters activated by a single TF, by two coactivators or in the presence of a competitive inhibitor. We find that target site search time and promoter occupancy differentially vary with the unspecific dissociation rate constant. Both parameters can be harmonized by adjusting the specific dissociation rate constant of the TF. However, while measured DNA residence times of various eukaryotic TFs correspond to a fast search time, the occupation frequencies of target sites are generally low. Cells might tolerate low target site occupancies as they enable timely gene regulation in response to a changing environment. | physics.bio-ph | physics | Title
Transcription factor target site search and gene regulation in a background of unspe-
cific binding sites
Authors
J. Hettich and J. C. M. Gebhardt*
Institute of Biophysics, Ulm University, Albert-Einstein-Allee 11, 89081 Ulm
*To whom correspondence should be addressed: [email protected]
Abstract
Response time and transcription level are vital parameters of gene regulation. They depend on how
fast transcription factors (TFs) find and how efficient they occupy their specific target sites. It is well
known that target site search is accelerated by TF binding to and sliding along unspecific DNA and that
unspecific associations alter the occupation frequency of a gene. However, whether target site search
time and occupation frequency can be optimized simultaneously is mostly unclear. We developed a
transparent and intuitively accessible state-based formalism to calculate search times to target sites
on and occupation frequencies of promoters of arbitrary state structure. Our formalism is based on
dissociation rate constants experimentally accessible in live cell experiments. To demonstrate our ap-
proach, we consider promoters activated by a single TF, by two coactivators or in the presence of a
competitive inhibitor. We find that target site search time and promoter occupancy differentially vary
with the unspecific dissociation rate constant. Both parameters can be harmonized by adjusting the
specific dissociation rate constant of the TF. However, while measured DNA residence times of various
eukaryotic TFs correspond to a fast search time, the occupation frequencies of target sites are generally
low. Cells might tolerate low target site occupancies as they enable timely gene regulation in response
to a changing environment.
Keywords
Facilitated diffusion, dimerization, competitive binding, master equation, multi-state model
1
Introduction
Gene regulation is mediated to large extend by transcription factors (TFs) interacting with specific tar-
get sites on DNA (van Hijum et al., 2009; Venters and Pugh, 2009). Frequently, TFs perform their regu-
latory function as homo- or heterodimers (Amoutzias et al., 2008), for example nuclear receptors
(Horwitz et al., 1996). Once bound, TFs recruit activating or repressing cofactors, chromatin remodelers
and the transcription machinery (Coulon et al., 2013). This recruiting activity and as a consequence the
regulatory function of TFs is more efficient the larger their probability to occupy the specific target site
(occupation frequency) is (Bain et al., 2012; Clauss et al., 2017).
The time a TF needs to find its specific target site in part determines the response time of a gene. It has
been proposed that the TF target site search is accelerated if the TF not only tests individual sites on
DNA by freely diffusing through the nucleoplasm (3D search), but in addition slides along DNA in an
unspecific binding mode (1D search)(Berg et al., 1981; Winter et al., 1981). Indeed, binding of proteins
to and sliding along unspecific DNA sequences is a common observation in reconstituted systems
(Blainey et al., 2006; Gorman et al., 2010; Kabata et al., 1993; Kim and Larson, 2007) and has been
observed in prokaryotes (Elf et al., 2007; Hammar et al., 2012). In mammalian cells, dissociation rate
constants from unspecific and specific binding sites are readily accessible by live cell experiments such
as single molecule imaging, fluorescence cross correlation or fluorescence recovery after photobleach-
ing (Gebhardt et al., 2013; Mazza et al., 2012), but direct observation of sliding has not yet been possi-
ble.
How various parameters such as the ratio of 1D to 3D search mechanisms, crowding, roadblocks or DNA
conformation influence the search time has been investigated comprehensively (Bauer and Metzler,
2012; de la Rosa et al., 2010; Gerland et al., 2002; Hu et al., 2006; Klenin et al., 2006; Koslover et al.,
2017; Krepel and Levy, 2016; Li et al., 2009; Shvets and Kolomeisky, 2016; Slutsky and Mirny, 2004;
Winter et al., 1981; Zabet and Adryan, 2013; Zabet et al., 2013). It is also known that unspecific binding
alters the occupation frequency of a gene (Gerland et al., 2002; Hippel et al., 1974; Schoech and Zabet,
2014; Slutsky and Mirny, 2004; Vonhippel and Berg, 1986). Yet it is unclear how both parameters com-
pare under variation of TF-DNA dissociation rate constants. How do TFs distribute over time between
their specific target site and the myriad of unspecific sites on DNA (Figure 1)? May a TF find its target
site quickly while at the same time occupying it efficiently to optimize gene regulation, or are both
requirements contradictory?
To answer these questions, we developed a simplified but powerful state-based formalism including
gene promoters of arbitrary state structure. TF target site search times are obtained accurately by solv-
ing the Kolmogorov equations for the time evolution to the steady state, and occupation frequencies
by considering the steady state distribution. Our formalism is based on dissociation rate constants and
thus readily enables calculation of target site search time and occupation frequency using experimen-
tally accessible values. We applied our approach to a simple promoter activated by a single TF and
validated our results by comparison with a microscopic search model (Mirny et al., 2009; Shvets and
Kolomeisky, 2016). Next, we expanded the activation mechanism by adding a cofactor dimerizing on
DNA or a competitor inhibiting specific binding of the TF. We found that both fast target site search
and high occupation frequency can be harmonized by adjusting the binding time to the specific target
2
site. Notably, experimentally obtained unspecific and specific dissociation rate constants of various TFs
indicate that many TFs indeed are optimized for fast target site search, but not for high occupation
frequencies of this site. Activating or competing cofactors allow modulating the optimal conditions,
enabling the cell to fine tune the response times and activities of genes in genetic networks.
Figure 1: General scheme of a TF binding to DNA. A specific binding site (e.g. a promoter) of arbitrary state struc-
ture is embedded in a pool of unspecific binding sites. Drawn is a simple promoter consisting of the target site
and surrounding unspecific DNA sequences. Depending on the respective binding times, the TF distributes be-
tween the nucleoplasm, unspecific sites and the promoter. If unspecific binding is favored, both the search time
and promoter occupancy decrease.
Materials and Methods
1. Mean number of free TFs
Single TF
The unspecific binding of the TF crowd is modelled by a chemical reaction scheme in which the rates
depend on the number of bound TFs and occupied binding sites (McGhee and Hippel, 1974). Let the
total number of unspecific binding sites be B. We assume that one binding site BS can only contain one
TF at a time. The chemical reaction for binding to unspecific sites is
1
If l TFs are bound to DNA, each TF can dissociate independently with the dissociation rate constant μ1.
The association rate constant of TFs is given by the product of free TFs N-l, the free binding sites B-l and
the diffusion limited arrival rate λ10. The stationary Kolmogorov equation for the probability pl to find l
bound TF is
3
TFBSBSTF+
2
This is a recurrence relation which can be solved for pl. This is done by calculating the initial value
The expectation value 𝑛 of unspecifically bound TFs is obtained by the derivative
3
4
Since this equation cannot be discussed analytically we next provide an approximation, which aims at
revealing the dependencies of the number of unspecifically bound TFs on the arrival and dissociation
rates as well as the number of unspecific binding sites. In order to find this approximation, we take the
average over Eq. 2 by multiplying with l and performing a summation with respect to l and obtain
For B, μ1/λ10>>N the approximation
5
6
is obtained. In this approximation the TF is diluted such that no saturation effects can occur. With this
approximation we obtain the free number of TFs
7
TF and competitor
For a second species C the binding to unspecific sites is dependent on the unspecific binding of the TF.
We again assume that a unspecific binding site can only be occupied by one molecule. The binding can
be written in terms of the chemical reaction
8
The stationary Kolmogorov equation is
4
()()()()()()() ),min(0for 0p where1110l11011BNMllplBlNplBlNpplllll==+−+−−−−−−+=−+p0-1=Nlaeèçöø÷l=0MåBlaeèçöø÷l!K0l where K0=l10m1 n=-K0ddK0ln(p0)2101)(nnNBBNn++−=BNn10111)(+=11011)(BNnNnf+=−=CBSBSCTFBSBSTF++
9
Here, association and dissociation of the TF are characterized by 10 and μ1, and of the competitor by
10 and 1, and the indices 1,2 with N and l denote TF and C, respectively. This recurrence relation can
be solved for p by calculating the initial value
10
where K= λ10/μ1 and H=Λ10/1. The expectation values 𝑛1, 𝑛2 of unspecifically bound TF and competitor
are obtained by the derivatives
11
Similar to the case of a single TF we provide an analytical solution for the regime in which TFs are di-
luted. We obtain two equations be averaging over l1 and l2
For B, μ1/λ10, ζ1/Λ10>>N1,N2 we obtain
This yields the mean free number of TFs and competitor C
12
13
14
5
()()()()()()()()()()()()()()),min(, ),min(0,for 0 where1111110222111211,2122,212210,1,21,12111,211110,,111212121212121212121BNMlBNMlllppllBlNpllBlNpplpllBlNpllBlNpplllllllllllllllllll===+−−+−−−−−−−+−+−−+−−−−−−−+=−+−+p0-1=N1l1aeèççöø÷÷l1,l2=0M1,M2åN2l2aeèççöø÷÷Bl1+l2aeèççöø÷÷(l1+l2)!Kl1Hl2n1=-KddKln(p0)n2=-HddHln(p0))()()()(212212222101211211111101nnnnnNnBBNnnnnnnNnBBNn+++−−=+++−−=BNnBNn10122101111 1+=+=11022,f1101,11 1BNnBNnf+=+=2. General formula for the target site search time and the occupation frequency of specific
target sites
We use a multistate model in order to model specific binding sites. On these sites two effects can occur:
unspecific binding and specific binding to the target site. In addition, the specific site can be empty with
respect to the TF. We introduce three index sets
15
E contains states in which no TF is bound, D contains states in which TFs are bound unspecifically, but
not to the specific target site and G contains states in which the specific target site is occupied. The
transition matrix T contains the transition rate constants between the states. The time evolution of the
system is described by the Kolmogorov equation for the binding probabilities p
16
The sum over all probabilities is normalized to 1.
We now calculate the target site search time. We define the search time as the time until any TF of the
crowd binds to the target site, starting from a situation in which the specific site is unoccupied. Since
we are interested in the first binding event we set all outgoing transitions of G to zero and thereby
obtain the matrix U. Next, we consider the flux of probabilities leaving the sets E,D
17
Where Pi denotes the probability to find the specific site in state i. In order to obtain this flux we need
to solve the Cauchy problem of the Kolmogorov equations
18
These initial conditions represent an empty specific site and therefore cover the whole target site
search which consists of 3D diffusion and transition to the target site from the unspecifically bound
state. The flux is the distribution of search times (Van Kampen, 1992). The expectation value of this
distribution is obtained by
19
20
By partial integration we obtain
where
6
},,{ },,{ },,{212121ggGddDeeE===ppT=−=DEiiptfdd==−00d()dddiiEDfttttptt=+=00diiiEDiEDiEDtpptQ
21
Thus, the mean search time can be calculated by the integral over time of the probabilities pi. These
can be obtained by integration over both sides of Eq. 18 which yields
22
We now calculate the probability p of the specific target site to be occupied (occupation frequency) by
solving the steady state Kolmogorov equation
23
for p. We obtain the occupation frequency by summing over all target states G.
Promoter activated by a single TF
We model the promoter (i.e. the specific site) with three states (Figure 5). The rate constants λ, λ1 and
λ2 correspond to the transitions form state 1 to 3, 1 to 2 and 2 to 3 respectively. The rate constants
denoted by μ represent the corresponding reverse reactions. The transition matrix of this three-state
model is
In order to represent the target site search problem, the states are divided into the sets
By solving Eq. 23 for p3 we obtain the occupation frequency
By setting transitions that leave the specific target sites to zero we obtain
By using Eq. 22 and the initial conditions
we obtain the target site search time
7
24
25
26
27
28
=0dtpQii )0()(QppU=−pT=0−−−−−−=22221111T}3{ }2{ }1{===GDEp=1+ml1+l1m1aeèçöø÷éëêùûú-1−−−−=000221111U)0,0,1()0( )1,0,0()(==pp
29
Promoter activated by two coactivating TFs
We assume that both TFs independently bind to two specific target sites within the promoter (i.e. the
specific site) with similar rate constants (Figure 2). We further assume that unspecific and specific sites
within the promoter can be occupied by two cofactors at the same time, while the unspecific sites of
the pool can only be occupied by one TF at a time. In order to represent the target site search problem,
the states are divided into the sets
30
The transition matrix of the model for two coactivating TFs (see Figure 2) is
31
We calculated the search time by numerically solving the system of equations Eq. 31.
The probability p for the target site to be occupied (occupation frequency) is calculated according to
Eq. 23 by detailed balance
32
with equilibrium constants from Figure 2.
8
++===++12112321 where ''Q}6{ }5,4,3,2{ }1{===GDE+++=+=+++=+++=+=−−−−−=212512421312121125212411321221112200000200000002000ddddddddddU2126)1(1KKKpp+++==Figure 2: Six-state model of a specific site (e.g. a promoter) that is active if two cofactors are bound to the specific
target site (red boxes). The unspecific site (white boxes) and the specific target site can contain two monomers
at the same time. The equilibrium constants are written in terms of the ratios bind/unbind with the rates Kl=λl/μl
l=1,2. The direct arrival at the target site with 3D diffusion of the protein is K=λ/μ
Promoter activated by a single TF subject to inhibitory competition
The competitor is assumed to inhibit binding of the TF within the promoter to unspecific sites or the
specific target site (Figure 3). Unspecific sites of the pool can also be bound by either the competitor or
the TF at a time. Since the unspecific and target sites can be empty or contain either a competitor or a
TF we obtain nine states. In order to represent the target site search problem, these states are divided
into the sets
33
Since the competitor presents a new kind of molecule with different association and dissociation rate
constants, we introduce ζ1 and ζ for the unspecific and specific dissociation rate constants. Similar to
the TF the competitor has arrival rates to unspecific and specific binding sites as well as a transition rate
from unspecific binding on the specific site to specific binding on the target site. We denote these rates
by Λ10, Λ0 and Λ2 respectively. The transition matrix of the model for the competitor (see Figure 3) is
9
}9,8,5{ }7,6,4,3,2{ }1{===GDE34
For simplicity we did not include the transitions to the set of states G in Eq. 34. The search time is
calculated by numerical matrix inversion of U.
The probability p for the target site to be occupied (occupation frequency) is calculated according to
Eq. 23 by detailed balance
35
with equilibrium constants from Figure 3.
Figure 3: Nine state model of the specific site (e.g. a promoter) for binding of a TF (green square) and a competitor
(blue star). On the promoter, the TF and the competitor can bind to the unspecific site (white boxes) and the
specific target site (red boxes). Both species occupy these sites exclusively. Since each site can have three states,
empty, competitor or TF bound, there are 32 = 9 states of the promoter. The equilibrium constants are written
in terms of the ratios bind/unbind with the rates Kl=λl/μl ; Hl=Λl/l l=1,2. The equilibrium constants H2,K2 are
chosen such that the law of detailed balance is fulfilled. The direct arrival at the specific target site with 3D
diffusion of the protein is K=λ/μ for the TF and H=Λ/ for the competitor.
10
+=+=+++=+++=+++=+++=−−−−−−=17162´142132112111716141321112211100000000000000ddddddddddddU)1(111985HKpppp++=++=−3. Model for target site search by explicit 1D diffusion
We here model the TF target site search by explicitly accounting for 1D diffusion events on unspecific
sites, interrupted by 3D diffusion until the specific target site is found (Bauer and Metzler, 2012; de la
Rosa et al., 2010; Gerland et al., 2002; Halford and Marko, 2004; Klenin et al., 2006; Koslover et al.,
2017; Krepel and Levy, 2016; Li et al., 2009; Mirny et al., 2009; Shvets and Kolomeisky, 2016; Slutsky
and Mirny, 2004; Vonhippel and Berg, 1989) (Figure 4). The association rate constant to any base pair
is λ0. We model the sliding of the TF by single base pair sliding steps with transition rate constant and
an average sliding time of μ1-1 (23). The transcription factor has one consensus DNA sequence to which
it can bind specifically. As in (Li et al., 2009; Shvets and Kolomeisky, 2016), we include static roadblock
molecules (crowders) in our model.
Figure 4: Model for TF target site search explicitly accounting for 1D and 3D diffusion based on a previous model
(23). State f corresponds to the free TF. State a corresponds to the TF bound to its specific target site. The TF may
arrive at any base pair of the genome by 3D diffusion and slides along unspecific DNA with sliding rate constant
. If bound unspecifically to the target site the TF may switch to the specific binding mode with transition rate
constant λ2,expl. The TF may also arrive in the specifically bound conformation directly after 3D diffusion with rate
constant 0. We accounted for crowding by periodically blocking sliding transitions between two sites every M
base pairs. With R crowders the total scanned genome length thus is L=R∙M.
We calculated the target site search time by applying our formalism of Materials and Methods, Section
2, to the initial value problem
36
Where f denotes the unbound state, a denotes the occupied specific target site and L denotes the length
of the scanned genome.
This leads to the result
11
E={f} D={-L/2,...,L/2} G={a}
37
where M is the number of base pairs of one unspecific binding site. In order to find the sliding rate
constant between base pairs, we consider continuous diffusion, where μ1→0, and omit crowding. In
this case the TF can slide along DNA without restrictions. We obtain for the target site search time
We compared this result to the formula
38
39
derived in (Mirny et al., 2009) which includes the 1D diffusion constant D measured experimentally in
(Elf et al., 2007; Gorman et al., 2010; Kim and Larson, 2007) and the length of one base pair lbp. The
number nD is the number of visited base pairs per sliding event. Both Eqs. 38 and 39 yield the same
results if we set
This result enables us to obtain the sliding rate constant
40
41
Using D=0.01 µm2s-2 (Gorman et al., 2010) and lbp=0.34 nm we obtain α=8.7∙104 s-1. The optimally fast
target site search time derived in (Mirny et al., 2009) requires a proportion of 1D diffusive search of
50%. Our model is consistent with this value in the limit case discussed above.
In Figure 6, our explicit diffusion model (Eq. 37) is plotted using the values N=100, B=104 and λ0=10−5 s−1
as motivated in the main text. The rate for transition to binding on the specific target site, λ2,expl=160 s-
1, was chosen such that the curves of the Diffusion model derived here agree with our new coarse-
grained model of Eq. 47.
12
121211111111121111.exp,21.exp,2.exp,2110−+−+=−−++=++++−+=−−qqqqqmmmmmLMMlll110/211+=L1102 where11DlnnLLbpDD=+=1122DlnbpD==2bplD=4. Application of the three-state model to eukaryotic TFs
According to our three-state model (Figure 5), a TF can dissociate from the specific target site using two
distinct pathways: direct dissociation from the specific target site with μ or transition to the intermedi-
ate, unspecifically bound state with μ2 followed by dissociation with μ1. Therefore, the experimentally
accessible specific dissociation rate constant represents an apparent rate constant determined by the
eigenvalues k of the Kolmogorov equations that include all dissociation pathways from the specific tar-
get site. In particular, the experimentally obtained specific dissociation rate constant does not directly
correspond to the rate constant μ. In order to apply our formalism to eukaryotic TFs, we therefore first
need to calculate the specific dissociation rate constant μ from the experimentally measured apparent
rate constant.
In order to calculate the eigenvalues k, we applied our formalism of Materials and Methods, Section 2,
to the initial value problem where states 2 and 3 are occupied and state 1 is empty.
We find the equation
for the eigenvalues.
42
Next, we obtain the specific dissociation rate constant µ using the measured rate constant k by elimi-
nating μ2 in Eq. 42 with the law of detailed balance and find
43
13
()()021212122=++++++−kk1121121−−+−+=kkMkResults
Target site search time and occupation frequency of a gene
Besides binding to their specific target site, TFs have an, albeit lower, affinity to any sequence on DNA
which gives rise to unspecific associations (Mustonen and Lassig, 2005; Slutsky et al., 2004; van Hijum
et al., 2009). Since dissociation rate constants from DNA can readily be measured in living cells, we set
out to develop a formalism for calculating the target site search time and occupation frequency of a
gene based on dissociation rate constants. We thus considered a specific site (e.g. a promoter) embed-
ded in a pool of unspecific binding sites (Figure 1) and translated this setting into a system of states
where the promoter is decoupled from the pool of unspecific sites (Figure 5).
We first calculated the mean number of free TFs considering the pool of unspecific binding sites at
equilibrium. Second, we solved the Kolmogorov equations for the specific site to obtain the time after
which the first TF of the crowd occupied the specific target site (target site search time) and the prob-
ability of the specific target site to be occupied (occupation frequency). We considered three different
promoter structures, a simple promoter bound by a single activating TF, a promoter activated by two
independently binding TFs with similar kinetic rate constants and a promoter to which an additional
inhibitory binding competitor can bind.
Figure 5: Three state model of TF-DNA binding. We consider a specific site for a TF (e.g. a promoter) including a
specific target site (red box) and an adjacent unspecific site (white box), as well as a pool of uncoupled unspecific
sites (left white box). The first state denotes free TFs, the second state denotes unspecifically bound TFs and the
third state denotes TFs bound to the specific target site. We ascribe each state an energy level, energy differences
are connected to the respective ratio of on- and off-rate constants by the law of detailed balance.
Promoter activated by a single TF
The simple promoter comprises the specific target site of the TF and surrounding unspecific DNA. Trans-
lated into our state-based formalism, the TF on the specific site (i.e. the promoter) has three states,
free, unspecifically bound and bound to the specific target site (Figure 5). The promoter is embedded in
a background of B unspecific binding sites from which no direct transition to the promoter can occur.
In our model, the TF can bind to its specific target site on the promoter either directly via 3D diffusion
14
with association rate constant λ, or by first binding to the surrounding unspecific DNA with association
rate constant λ1 and a subsequent transition with rate constant λ2. λ2 is a cumulative rate including a
potential conformational change associated with binding to the target site (Slutsky and Mirny, 2004)
and a finite probability to find the target site when sliding on nearby unspecific sequences (Hammar et
al., 2012). For the size of unspecific sites we chose a fixed number of 40 bp, corresponding to experi-
mental findings (Gorman et al., 2010; Hammar et al., 2012). In vivo, the size of unspecific sites is limited
by the presence of other DNA-bound molecules, not by the time a TF spends sliding on accessible DNA
(Li et al., 2009; Shvets and Kolomeisky, 2016), justifying a fixed size independent of µ1 .
The association rate constants λ1 and λ to any unspecific site or the specific target site depend on the
number of free TFs, nf and the arrival rates of a single TF λ10 and λ0 on unspecific and specific sites. This
number in turn depends on the dissociation rate constant of unspecific sites, µ1, according to
44
Here, we neglect a dependence on the dissociation rate constant of the specific target site since only
one such site exists, which is not occupied during target site search. λ0 and λ10 denote the diffusion
limited arrival rate constants.
Association rate constants and dissociation rate constants are coupled by the law of detailed balance
45
We first calculated the number of free TFs nf (Material and Methods, Section 1). If the total number of
TFs N is much smaller than B and the ratio μ1/λ10, we find for nf
46
This equation shows that the number of free TFs significantly decreases if the number of unspecific sites
is large and unspecific binding is favoured by a large ratio λ10/µ1. For the following discussion we used
the exact result for nf (Eq. 4 derived in Materials and Methods, Section 1) and evaluated it numerically.
Next, we calculated the target site search time of the gene. We assumed an equilibrium distribution
of free and unspecifically bound TFs in the pool of unspecific binding sites, given an unoccupied pro-
moter, and used the Kolmogorov equations to calculate the time until the first TF bound the target site
(Materials and Methods, Section 2). Using the number of free TFs, we found for the target site search
time (Figure 6)
47
where λ2 is the transition from unspecific binding on the specific site to binding to the specific target
site. The search time exhibits a minimum corresponding to an optimally fast target search process, as
for small µ1 TFs are drawn towards unspecific DNA and the free pool of searching TFs becomes minimal
15
l1m1()=nfm1()l10 lm1()=nfm1()l0l10m1×l2m2=l0m11011)(BNnf+=1211221 where ''++==++while for large µ1 the benefit of 1D diffusional scanning of DNA is negligible. This result is consistent
with previous studies of the TF target site search problem that explicitly included 1D diffusion (Berg et
al., 1981; Halford and Marko, 2004; Mirny et al., 2009; Winter et al., 1981).
Figure 6: : Comparison of the target site search time of the gene as function of the unspecific dissociation rate
constant µ1 between a promoter activated by a single TF (TF, red line), a promoter activated by a dimer (Dimer
µ=10-3 s-1, solid blue line), activated by a dimer with lower specific dissociation rate constant (Dimer µ=10-4 s-1,
dashed blue line) and when an inhibitory competitor of the TF is present (Competitor, yellow line). In addition,
the TF target site search time calculated from a model explicitly accounting for 1D diffusion is indicated (Diffusion
model, dashed black line). Simulation parameters are: N=100, C =100, B=104, λ10=Λ10=4·10−4 s−1, λ0=Λ0=10−5 s−1,
λ2=Λ2=4 s−1, ζ1=1 s−1, ζ=10−3 s−1.
To validate our simplified calculation of the target site search time, we compared our result with the
one from a microscopic model explicitly treating 1D diffusion in presence of static roadblocks that limit
the sliding distance of TFs (Li et al., 2009; Shvets and Kolomeisky, 2016) (Materials and Methods Section
3, Figure 4). Both models agree very well (Figure 6). Since our simplified state based approach allows
application to experimentally measured TF rate constants and enables straightforward incorporation
of more complex promoter structures, we used it hereafter.
We calculated the probability p of the specific target site to be occupied (occupation frequency) of the
gene by eliminating λ2/μ2 with Eq. 45 and solved the Kolmogorov equations (Materials and Methods,
Section 2). We found for the occupation frequency (Figure 7)
16
48
This parameter increases monotonically with µ1 and asymptotically approaches a maximum deter-
mined by the equilibrium constant of specific binding, since µ1 directly determines the number of free
TFs and thus the association rate constant to specific target sites (Eq. 44).
Figure 7: Comparison of occupation frequency p as function of the unspecific dissociation rate constant µ1 be-
tween a promoter activated by a single TF (TF, red line), a promoter activated by a dimer formed on DNA (Dimer
on DNA, blue line) and by a dimer formed in the nucleoplasm (Dimer in nucleoplasm, dashed blue line), when an
inhibitory competitor of the TF is present (Competitor ζ1=1 s−1, yellow line) and when the competitor has a larger
unspecific dissociation rate constant (Competitor ζ1=25 s−1, dashed yellow line). Simulation parameters are:
N=100, C =100, B=104, λ10=Λ10=4·10−4 s−1, λ0=Λ0=10−5 s−1, λ2=Λ2=4 s−1, μ=10−1 s−1, ζ=10−3 s−1.
When comparing Figures 6 and 7, it becomes obvious that the unspecific dissociation rate constant µ1
at which the target site search time is minimal does not necessarily correspond to a high occupation
frequency, indicating that target site search time and occupation frequency might not simultaneously
be optimized for a given TF. However, since the occupation frequency depends on the specific dissoci-
ation rate constant while the target site search time does not, both parameters might be optimized by
adjusting the specific dissociation rate constant.
Promoter activated by two coactivating TFs
To model gene activation by two independently binding TFs of similar type, the promoter comprises
two specific target sites, one for each TF, surrounded by unspecific DNA. Both TFs are assumed to bind
17
p=1+ml1+l1m1aeèçöø÷éëêùûú-1independently with similar rate constants but activate the gene only if bound specifically at the same
time as homo- or heterodimers. Translated into our state-based formalism, the TFs on the specific site
have six states (Materials and Methods Section 2, Figure 2). We deliberately omitted binding interac-
tions between both TFs (Geisel and Gerland, 2011), to focus on the contributions of DNA binding to
target site search and occupation frequency and preserve the comparison to a single TF.
To calculate the target site search time of the promoter activated by two TFs, we solved the Kolmogorov
equations of the dimeric system numerically (Figure 6 and Materials and Methods, Section 2). For the
occupation frequency, we compared dimerization exclusively on DNA with dimerization already in so-
lution (Figure 7).
While for a single TF the target site search time is independent of the specific dissociation rate constant,
two independently binding coactivating TFs simultaneously find the specific target site the faster, the
smaller the specific dissociation rate constant is. This reflects the fact that both factors have to wait for
each other on the specific target site. Accordingly, the probability of simultaneous occupation of the
target site is very low for two independently binding TFs. This occupation frequency increases if dimer-
ization is allowed to occur already in the nucleoplasm.
Promoter activated by a single TF subject to inhibitory competition
To model gene activation in the presence of an inhibitory binding competitor, we introduced a second
species that is able to bind to the same unspecific and specific binding sites as the TF, but with own
dissociation rate constants. Once occupied by the competitor, a binding site is sterically blocked for
binding of the TF. Translated into our state-based formalism, the system has nine states (Materials and
Methods Section 2, Figure 3).
We again numerically solved the Kolmogorov equations of the system with binding competitor to ob-
tain the target site search time of the gene. The target site search time of the TF increases significantly
if the unspecific dissociation rate constant ζ1 of the competitor increases (Figure 8a). In this case more
competitor is available to block the target site. Consequently, the occupation frequency of the gene
drops under these conditions (Figure 8b). If the occupation of the specific target site by the competitor
is modulated with the specific dissociation rate constant ζ of the competitor, a small value for ζ, i.e. a
high occupation, leads to a similar effect (Figure 8c and d). These findings agree well with (Zabet and
Adryan, 2013), where an increased coverage of DNA by mobile competitors and with that of the specific
binding site was achieved by increasing competitor abundance.
18
Figure 8: Influence of an inhibitory binding competitor. (A) Target site search time of the gene and (B) proba-
bility p of the promoter to be occupied by a TF (occupation frequency) as function of the unspecific dissociation
rate constant 1 of the competitor. (C) and (D) p as function of the specific dissociation rate constant of the
competitor. Simulation parameters are: N=100, C =100, B=104, λ10=Λ10=4·10−4 s−1, λ0=Λ0=10−5 s−1, λ2=Λ2=4 s−1,
μ=10−1 s−1, ζ=10−3 s−1.
TFs are optimized for fast target site search but not for high occupation frequency
Since both target site search time and occupation frequency depend on the unspecific dissociation rate
constant µ1, but only the occupation frequency depends on the specific dissociation rate constant µ,
we tested under which conditions target site search and occupation frequency assumed their respec-
tive optima, fast target site search and high occupation frequency (Figure 9). To compare both quanti-
ties, we chose parameter values according to the physiological conditions in an eukaryotic cell. We
varied the unspecific dissociation rate constant µ1 in the interval 100 s-1 to 103 s-1 (Ball et al., 2016;
Caccianini et al., 2015; Elf et al., 2007; Morisaki et al., 2014) and the specific dissociation rate constant
µ from 100 s-1 to 10-4 s-1 (Agarwal et al., 2017; Caccianini et al., 2015; Chen et al., 2014; Gebhardt et al.,
2013; Hammar et al., 2014; Mazza et al., 2012; Speil et al., 2011; Sugo et al., 2015). The arrival rate
constant is given by the 3D diffusion limit under physiological salt concentrations, λ0=106 (Ms)-1 (Mirny
et al., 2009), which results in λ0=10-5 s-1 in a typical cell volume of 100μm3. The arrival rate constant to
the unspecific site scales with the size of unspecific sites of 40bp to λ10=4·10-4 s-1. We adjusted the
transition rate constant λ2 =4 s-1 in accordance with a low probability to bind the specific target site
19
when scanning nearby unspecific DNA (Hammar et al., 2012). In this case the rate limiting step of spe-
cific DNA-binding is the 3D diffusion to the specific site. Finally, we assumed that the promoter is em-
bedded in a background of 4·105 bp of unspecific DNA, resulting in B=104 unspecific binding sites. This
value revealed a fraction of DNA-associated TFs of 20%-40% (Materials and Methods Section 1, Eq. 6)
reported by experiments (Chen et al., 2014; Gebhardt et al., 2013; Mazza et al., 2012). This fraction
corresponds to the fraction of time the TF spends in 1D search. We found that a common optimum of
target site search time and occupation frequency is only achieved for large unspecific and small specific
dissociation rate constants (Figure 9).
We next applied our model to various eukaryotic TFs. We calculated their occupation frequencies using
published values of unspecific and specific dissociation rate constants measured in vivo, and plotted
these onto the theoretically derived occupation frequency (Figure 9 and Table 1). It is important to note
that experimentally accessible dissociation rate constants are apparent rate constants that first have to
be converted to the specific dissociation rate constant of the three-state model (Materials and Meth-
ods, Section 4). To be able to compare the different TFs, we assumed that dimeric TFs such as GR already
dimerized in solution, and considered a ratio of 100 TFs per specific target site, which is a good approx-
imation for many eukaryotic TFs binding to hundreds of specific target sequences (Geiger et al., 2012).
However, we note that calculation of the target site search time requires knowledge about the number
of TFs present in the cell and the ratio of specific to unspecific binding sites, which will be specific to
each TF. We found that all considered TFs are optimized for a fast target site search time. In contrast,
the probability of most TFs to occupy their specific target site is low. Only TFs with very small specific
dissociation rate constants such as CTCF have a high occupation frequency.
Figure 9: Occupation frequency of the promoter by a TF as function of the unspecific (µ1) and the specific (µ)
dissociation rate constants. The values µ1 of minimal (red line) and 1.5 times the minimal target site search time
(white lines) are indicated. The green line at µ=10-1 s-1 indicates the cross section of the occupation frequency
displayed in Figure 7. Symbols correspond to occupation frequencies of various TFs calculated from unspecific
and specific dissociation rate constants measured by fluorescence microscopy in live cells. Simulation parameters
are: N=100, B=104, λ10 =4·10−4 s−1, λ0 =10−5 s−1, λ2= 4 s−1.
20
μ1 (s-1)
1.3
5.6
k (s-1)
0.08
0.003
0.2
0.3
0.4
0.5
TF
Sox2
LacI (mamma-
lian cell)
CREB
P53
Stat1
GR
CTCF
2.7
20
50
6.7
5
μ (s-1)
0.003
0.0002
0.009
0.04
0.1
0.03
source
(Chen et al., 2014)
(Caccianini et al., 2015;
Hammar et al., 2014)
(Sugo et al., 2015)
(Mazza et al., 2012)
(Speil et al., 2011)
(Gebhardt et al., 2013;
Groeneweg et al., 2014)
(Agarwal et al., 2017)
0.001
0.00005
Table 1: Unspecific (µ1) and apparent specific (k) dissociation rate constants of transcription factors measured in
live cells. The apparent dissociation rate constant k is converted to the specific dissociation rate constant of the
three-state model,μ, using Eq. 43.
21
Discussion
We have introduced a treatment of TF target site search that is based on TF-DNA dissociation rate con-
stants accessible with live cell measurements and thus readily applicable to experiments. To achieve
this, we replaced explicit 1D diffusion of the TF with its dissociation rate constant from unspecific sites.
We have shown that this replacement is able to model the search process in presence of explicit diffu-
sion and static roadblock molecules or crowding molecules (Koslover et al., 2017; Krepel and Levy, 2016;
Li et al., 2009; Morelli et al., 2011; Shvets and Kolomeisky, 2016) to good approximation. Our formalism
thus yields very transparent and intuitively accessible analytical solutions for the target site search
problem.
When applying our model to experimentally measured dissociation rate constants of eukaryotic TFs,
we find that TFs indeed operate at the optimum of a fast target site search, but high occupation fre-
quencies are not necessarily achieved in cells. This has also been observed for Drosophila TFs (Zabet
and Adryan, 2015). While the low occupation frequency could in principle be counteracted by a high
specific dissociation rate constant, such a strategy does not seem to occur for most TFs. Another way
to increase the occupation frequency is to increase the total number of TFs. This effect becomes par-
ticularly important if the TF number is comparable to the number of unspecific binding sites. A possible
reason for suboptimal occupation frequency could arise from the stability paradox (Gerland et al., 2002;
Slutsky and Mirny, 2004), which states that the energy gap between unspecific and specific binding
cannot be arbitrarily high and a compromise between both values might need to be attained. On the
other hand, these particular choices of specific dissociation rate constants may also be advantageous
for the functioning of TFs. Most eukaryotic TFs act not only on one promoter but regulate hundreds of
genes and contribute to different signalling cascades. In such a genetic network short specific binding
times ensure that always a large fraction of molecules is available in solution and the time to find a
specific target site is kept low. This enables fast reactions to changing signalling environments. In con-
trast, the long specific binding time measured for the chromatin architecture protein CTCF (Agarwal et
al., 2017) leads to a high fraction of occupied target sites and a considerable reduction of free protein.
Thus, the target site search time will be high once a CTCF molecule dissociated and CTCF-mediated
chromatin loops will only reform slowly.
Optimizing for fast target site search and high occupation frequency may not be the only constraint in
the evolution of dissociation rate constants. Given the complexity of gene regulation, target site search
might not be the rate limiting step. Additionally, other requirements such as fast reaction to signals or
a kinetic separation of monomer and dimer function as discussed below may shape TF dissociation rate
constants.
In addition to being easily applicable to experimentally accessible data, our target search formalism
simplifies the search problem by mathematically decoupling the discussion of TF binding-modalities at
the specific site (e.g. different promoter structures) from unspecific binding to the pool of unspecific
binding sites. Accordingly, our initial condition for the target site search is equilibrium between free
and unspecifically bound TFs, and we calculated the time after which the first TF of the crowd binds to
the specific target site. This is in contrast to the common approach considering the full search pathway
of a single TF. However, since the transient process toward the initial equilibrium converges with a rate
22
of B∙λ1, which is much faster than the target site search time, our assumption for the initial condition
barely influences the search time. Thus, while being coarse-grained, our model is still general and pow-
erful. It can be further extended to include the topology of DNA and crowding effects in the nucleo-
plasm (Hu et al., 2006; Kamar et al., 2017; Morelli et al., 2011), by appropriate changes to the arrival
and dissociation rate constants. In particular, protein-protein interactions could easily be incorporated
using additional states. Our model may in addition accelerate the simulation of a background of unspe-
cific binding sites, by replacing modelling of myriads of explicit 1D sliding states by the unspecific disso-
ciation rate. A similar technique was applied to scale and accelerate the simulation of the entire facili-
tated diffusion process (Zabet, 2012; Zabet and Adryan, 2012).
A state-based formalism was previously used in (Schoech and Zabet, 2014) to investigate the buffering
effect of facilitated diffusion on gene expression noise. This state-based model combines 1D and 3D
diffusion in a single association rate constant to the specific target site. Rebinding events to the target
site by sliding are explicitly considered by an additional state and shown to be equivalent to an effective
dissociation rate constant. In our approach, we used the state-based formalism to model the search
time of the TF for its specific target site in more detail by including a state for unspecific binding.
The non-monotone dependency of the target site search time on the duration of 1D sliding, respectively
the unspecific dissociation rate constant in our approach, has been discussed comprehensively in pre-
vious studies (Bauer and Metzler, 2012; de la Rosa et al., 2010; Gerland et al., 2002; Klenin et al., 2006;
Mirny et al., 2009; Vonhippel and Berg, 1989). Here, we focused on how the search process adapts to
more complex promoters that involve coactivators and inhibitory competitors. The microscopic diffu-
sion processes constrained by position and overlap of binding sites for competitors and TFs were pre-
viously motivated and investigated (Ezer et al., 2014).
We considered the limit cases of exclusive dimerization in the nucleoplasm and exclusive dimerization
on the target site in absence of a stabilizing protein-protein interaction (Geisel and Gerland, 2011). This
allows direct comparison of search times and occupation frequencies of two coactivating TFs with a
single TF. We found that two independently binding TFs prolong the target site search time significantly.
Additionally, the unspecific dissociation rate constant μ1 for an optimal target site search time in-
creases. This increase is accompanied by a decrease of the transition rate constant μ2 from a specifically
to an unspecifically bound TF due to the law of detailed balance, meaning that the two TFs have to wait
for each other on the specific target site to achieve fastest gene response times. Alternatively, dimeric
TFs may wait for each other by increasing the specific dissociation rate constant µ. Whether dimeric TFs
generally exhibit larger dissociation rate constants from unspecific DNA or smaller dissociation rate
constants from specific target sites compared to solely activating TFs is subject to future studies.
For inducible dimers capable to dimerize in solution, such as nuclear receptors (Horwitz et al., 1996) ,
the non-linear increase of target site search time at low dissociation rate constants may lead to a kinetic
separation of monomer and dimer function. While uninduced monomers in principle could simultane-
ously arrive at the promoter, potentially evoking background activation, such a search mechanism is
much slower than the one for dimer molecules that effectively bind as one TF. This turns background
activation of uninduced monomeric molecules inefficient. The occupation frequency of the promoter
activated by two TFs drops significantly compared to the case of a single TF. The lower gene activation
23
capability arises due to the more complex activation mechanism and could be accounted for by increas-
ing the monomer concentration or by allowing dimerization not only on DNA but already in the nucle-
oplasm. This constitutes an additional advantage for solution-based dimerization and sheds new light
on the question whether for example nuclear receptors bind to DNA as monomers or dimers (Ong et
al., 2010).
We have discussed the influence of a competitor that inhibits binding to the specific target site. The
modulatory effect of an inhibitory competitor on the target site search time and occupation frequency
of a TF can be mainly ascribed to blocking of the specific site rather than preventing the TF from scan-
ning unspecific DNA sequences common to all DNA binding proteins. Thus, an inhibitory competitor will
indeed only alter the activity of a certain gene without changing the search times or occupation fre-
quencies of other genes (Clauss et al., 2017). In contrast, an unspecific competitor will mainly affect TF
occupation at unspecific sites, turning occupation of the target site more susceptible to TF abundance
(Zabet et al., 2013).
Acknowledgements
We thank Matthias Reisser for helpful discussions. This work was supported by the German Research
Foundation [GE 2631/1-1 to J.C.M.G.] and the European Research Council (ERC) under the European
Union's Horizon 2020 Research and Innovation Programme [637987 ChromArch to J.C.M.G.].
Author Contributions
J.H. and J.C.M.G. designed the study; J.H. performed calculations; J.H. and J.C.M.G. wrote the manu-
script.
References
Agarwal, H., Reisser, M., Wortmann, C., Gebhardt, J. C. M., 2017. Direct Observation of Cell-Cycle-Dependent
Interactions between CTCF and Chromatin. Biophysical Journal 112, 2051-2055,
doi:10.1016/j.bpj.2017.04.018.
Amoutzias, G. D., Robertson, D. L., de Peer, Y. V., Oliver, S. G., 2008. Choose your partners: dimerization in
eukaryotic transcription factors. Trends in Biochemical Sciences 33, 220-229,
doi:10.1016/j.tibs.2008.02.002.
Bain, D. L., Yang, Q., Connaghan, K. D., Robblee, J. P., Miura, M. T., Degala, G. D., Lambert, J. R., Maluf, N. K.,
2012. Glucocorticoid Receptor-DNA Interactions: Binding Energetics Are the Primary Determinant of
Sequence-Specific Transcriptional Activity. Journal of Molecular Biology 422, 18-32,
doi:10.1016/j.jmb.2012.06.005.
Ball, D. A., Mehta, G. D., Salomon-Kent, R., Mazza, D., Morisaki, T., Mueller, F., McNally, J. G., Karpova, T. S.,
2016. Single molecule tracking of Ace1p in Saccharomyces cerevisiae defines a characteristic residence
time for non-specific interactions of transcription factors with chromatin. Nucleic Acids Research 44,
e160, doi:10.1093/nar/gkw744.
Bauer, M., Metzler, R., 2012. Generalized Facilitated Diffusion Model for DNA-Binding Proteins with Search and
Recognition States. Biophysical Journal 102, 2321-2330, doi:10.1016/j.bpj.2012.04.008.
24
Berg, O. G., Winter, R. B., Vonhippel, P. H., 1981. Diffusion-Driven Mechanisms of Protein Translocation on
Nucleic-Acids .1. Models and Theory. Biochemistry 20, 6929-6948, doi:10.1021/bi00527a028.
Blainey, P. C., van Oijent, A. M., Banerjee, A., Verdine, G. L., Xie, X. S., 2006. A base-excision DNA-repair protein
finds intrahelical lesion bases by fast sliding in contact with DNA. Proceedings of the National Academy
of Sciences of the United States of America 103, 5752-5757, doi:10.1073/pnas.0509723103.
Caccianini, L., Normanno, D., Izeddin, I., Dahan, M., 2015. Single molecule study of non-specific binding kinetics
of Lacl in mammalian cells. Faraday Discussions 184, 393-400, doi:10.1039/c5fd00112a.
Chen, J. J., Zhang, Z. J., Li, L., Chen, B. C., Revyakin, A., Hajj, B., Legant, W., Dahan, M., Lionnet, T., Betzig, E.,
Tjian, R., Liu, Z., 2014. Single-Molecule Dynamics of Enhanceosome Assembly in Embryonic Stem Cells.
Cell 156, 1274-1285, doi:10.1016/j.cell.2014.01.062.
Clauss, K., Popp, A. P., Schulze, L., Hettich, J., Reisser, M., Escoter Torres, L., Uhlenhaut, N. H., Gebhardt, J. C. M.,
2017. DNA residence time is a regulatory factor of transcription repression. Nucleic Acids Research 45,
11121–11130.
Coulon, A., Chow, C. C., Singer, R. H., Larson, D. R., 2013. Eukaryotic transcriptional dynamics: from single
molecules to cell populations. Nature Reviews Genetics 14, 572-584, doi:10.1038/nrg3484.
de la Rosa, M. A. D., Koslover, E. F., Mulligan, P. J., Spakowitz, A. J., 2010. Dynamic Strategies for Target-Site
Search by DNA-Binding Proteins. Biophysical Journal 98, 2943-2953, doi:10.1016/j.bpj.2010.02.055.
Elf, J., Li, G. W., Xie, X. S., 2007. Probing transcription factor dynamics at the single-molecule level in a living
cell. Science 316, 1191-1194, doi:10.1126/science.1141967.
Ezer, D., Zabet, N. R., Adryan, B., 2014. Physical constraints determine the logic of bacterial promoter
architectures. Nucleic Acids Research 42, 4196-4207, doi:10.1093/nar/gku078.
Gebhardt, J. C. M., Suter, D. M., Roy, R., Zhao, Z. Q. W., Chapman, A. R., Basu, S., Maniatis, T., Xie, X. S., 2013.
Single-molecule imaging of transcription factor binding to DNA in live mammalian cells. Nature
Methods 10, 421-+, doi:10.1038/nmeth.2411.
Geiger, T., Wehner, A., Schaab, C., Cox, J., Mann, M., 2012. Comparative Proteomic Analysis of Eleven Common
Cell Lines Reveals Ubiquitous but Varying Expression of Most Proteins. Molecular & Cellular Proteomics
11, M111-014050, doi:10.1074/mcp.M111.014050.
Geisel, N., Gerland, U., 2011. Physical Limits on Cooperative Protein-DNA Binding and the Kinetics of
Combinatorial Transcription Regulation. Biophysical Journal 101, 1569-1579,
doi:10.1016/j.bpj.2011.08.041.
Gerland, U., Moroz, J. D., Hwa, T., 2002. Physical constraints and functional characteristics of transcription
factor-DNA interaction. Proceedings of the National Academy of Sciences of the United States of
America 99, 12015-12020, doi:10.1073/pnas.192693599.
Gorman, J., Plys, A. J., Visnapuu, M. L., Alani, E., Greene, E. C., 2010. Visualizing one-dimensional diffusion of
eukaryotic DNA repair factors along a chromatin lattice. Nature Structural & Molecular Biology 17, 932-
U37, doi:10.1038/nsmb.1858.
Groeneweg, F. L., van Royen, M. E., Fenz, S., Keizer, V. I. P., Geverts, B., Prins, J., de Kloet, E. R., Houtsmuller, A.
B., Schmidt, T. S., Schaaf, M. J. M., 2014. Quantitation of Glucocorticoid Receptor DNA-Binding
Dynamics by Single-Molecule Microscopy and FRAP. Plos One 9, doi:10.1371/journal.pone.0090532.
Halford, S. E., Marko, J. F., 2004. How do site-specific DNA-binding proteins find their targets? Nucleic Acids
Research 32, 3040-3052, doi:10.1093/nar/gkh624.
Hammar, P., Leroy, P., Mahmutovic, A., Marklund, E. G., Berg, O. G., Elf, J., 2012. The lac Repressor Displays
Facilitated Diffusion in Living Cells. Science 336, 1595-1598, doi:10.1126/science.1221648.
Hammar, P., Wallden, M., Fange, D., Persson, F., Baltekin, O., Ullman, G., Leroy, P., Elf, J., 2014. Direct
measurement of transcription factor dissociation excludes a simple operator occupancy model for gene
regulation. Nature Genetics 46, 405-+, doi:10.1038/ng.2905.
Hippel, P. H. V., Revzin, A., Gross, C. A., Wang, A. C., 1974. Non-specific DNA Binding of Genome Regulating
Proteins as a Biological Control Mechanism: 1. The lac Operon: Equilibrium Aspects. Proceedings of the
National Academy of Sciences 71, 4808-4812.
25
Horwitz, K. B., Jackson, T. A., Rain, D. L., Richer, J. K., Takimoto, G. S., Tung, L., 1996. Nuclear receptor
coactivators and corepressors. Molecular Endocrinology 10, 1167-1177, doi:10.1210/me.10.10.1167.
Hu, T., Grosberg, A. Y., Shklovskii, B. I., 2006. How proteins search for their specific sites on DNA: The role of
DNA conformation. Biophysical Journal 90, 2731-2744, doi:10.1529/biophysj.105.078162.
Kabata, H., Kurosawa, O., Arai, I., Washizu, M., Margarson, S. A., Glass, R. E., Shimamoto, N., 1993. Visualization
of Single Molecules of RNA-Polymerase Sliding Along DNA. Science 262, 1561-1563,
doi:10.1126/science.8248804.
Kamar, R. I., Banigan, E. J., Erbas, A., Giuntoli, R. D., de la Cruz, M. O., Johnson, R. C., Marko, J. F., 2017.
Facilitated dissociation of transcription factors from single DNA binding sites. Proceedings of the
National Academy of Sciences of the United States of America 114, E3251-E3257,
doi:10.1073/pnas.1701884114.
Kim, J. H., Larson, R. G., 2007. Single-molecule analysis of 1D diffusion and transcription elongation of T7 RNA
polymerase along individual stretched DNA molecules. Nucleic Acids Research 35, 3848-3858,
doi:10.1093/nar/gkm332.
Klenin, K. V., Merlitz, H., Langowski, J., Wu, C. X., 2006. Facilitated diffusion of DNA-binding proteins. Physical
Review Letters 96, 018104, doi:10.1103/PhysRevLett.96.018104.
Koslover, E. F., de la Rosa, M. D., Spakowitz, A. J., 2017. Crowding and hopping in a protein's diffusive transport
on DNA. Journal of Physics a-Mathematical and Theoretical 50, 074005, doi:10.1088/1751-
8121/aa53ee.
Krepel, D., Levy, Y., 2016. Protein diffusion along DNA: on the effect of roadblocks and crowders. Journal of
Physics a-Mathematical and Theoretical 49, 494003, doi:10.1088/1751-8113/49/49/494003.
Li, G. W., Berg, O. G., Elf, J., 2009. Effects of macromolecular crowding and DNA looping on gene regulation
kinetics. Nature Physics 5, 294-297, doi:10.1038/nphys1222.
Mazza, D., Abernathy, A., Golob, N., Morisaki, T., McNally, J. G., 2012. A benchmark for chromatin binding
measurements in live cells. Nucleic Acids Research 40, e119, doi:10.1093/nar/gks701.
McGhee, J. D., Hippel, P. H. V., 1974. Theoretical Aspects of DNA-Protein Interactions - cooperative and Non-
Cooperative Binding of Large Ligands to a One-Dimensional Homogeneous Lattice. Journal of Molecular
Biology 86, 469-489, doi:10.1016/0022-2836(74)90031-x.
Mirny, L., Slutsky, M., Wunderlich, Z., Tafvizi, A., Leith, J., Kosmrlj, A., 2009. How a protein searches for its site
on DNA: the mechanism of facilitated diffusion. Journal of Physics a-Mathematical and Theoretical 42,
434013, doi:10.1088/1751-8113/42/43/434013.
Morelli, M. J., Allen, R. J., ten Wolde, P. R., 2011. Effects of Macromolecular Crowding on Genetic Networks.
Biophysical Journal 101, 2882-2891, doi:10.1016/j.bpj.2011.10.053.
Morisaki, T., Muller, W. G., Golob, N., Mazza, D., McNally, J. G., 2014. Single-molecule analysis of transcription
factor binding at transcription sites in live cells. Nature Communications 5, 4456,
doi:10.1038/ncomms5456.
Mustonen, V., Lassig, M., 2005. Evolutionary population genetics of promoters: Predicting binding sites and
functional phylogenies. Proceedings of the National Academy of Sciences of the United States of
America 102, 15936-15941, doi:10.1073/pnas.0505537102.
Ong, K. M., Blackford, J. A., Kagan, B. L., Simons, S. S., Chow, C. C., 2010. A theoretical framework for gene
induction and experimental comparisons. Proceedings of the National Academy of Sciences of the
United States of America 107, 7107-7112, doi:10.1073/pnas.0911095107.
Schoech, A. P., Zabet, N. R., 2014. Facilitated diffusion buffers noise in gene expression. Physical Review E 90,
doi:10.1103/PhysRevE.90.032701.
Shvets, A. A., Kolomeisky, A. B., 2016. Crowding on DNA in Protein Search for Targets. Journal of Physical
Chemistry Letters 7, 2502-2506, doi:10.1021/acs.jpclett.6b00905.
Slutsky, M., Mirny, L. A., 2004. Kinetics of protein-DNA interaction: Facilitated target location in sequence-
dependent potential. Biophysical Journal 87, 4021-4035, doi:10.1529/biophysj.104.050765.
Slutsky, M., Kardar, M., Mirny, L. A., 2004. Diffusion in correlated random potentials, with applications to DNA.
Physical Review E 69, 11, doi:10.1103/PhysRevE.69.061903.
26
Speil, J., Baumgart, E., Siebrasse, J. P., Veith, R., Vinkemeier, U., Kubitscheck, U., 2011. Activated STAT1
Transcription Factors Conduct Distinct Saltatory Movements in the Cell Nucleus. Biophysical Journal
101, 2592-2600, doi:10.1016/j.bpj.2011.10.006.
Sugo, N., Morimatsu, M., Arai, Y., Kousoku, Y., Ohkuni, A., Nomura, T., Yanagida, T., Yamamoto, N., 2015.
Single-Molecule Imaging Reveals Dynamics of CREB Transcription Factor Bound to Its Target Sequence.
Scientific Reports 5, 9, doi:10.1038/srep10662.
van Hijum, S., Medema, M. H., Kuipers, O. P., 2009. Mechanisms and Evolution of Control Logic in Prokaryotic
Transcriptional Regulation. Microbiology and Molecular Biology Reviews 73, 481-+,
doi:10.1128/mmbr.00037-08.
Van Kampen, N. G., 1992. Stochastic processes in physics and chemistry. Elsevier.
Venters, B. J., Pugh, B. F., 2009. How eukaryotic genes are transcribed. Critical Reviews in Biochemistry and
Molecular Biology 44, 117-141, doi:10.1080/10409230902858785.
Vonhippel, P. H., Berg, O. G., 1986. On the specificity of DNA_Protein Interactions. Proceedings of the National
Academy of Sciences of the United States of America 83, 1608-1612.
Vonhippel, P. H., Berg, O. G., 1989. Facilitated Target Location in Biological-Systems. Journal of Biological
Chemistry 264, 675-678.
Winter, R. B., Berg, O. G., Vonhippel, P. H., 1981. Diffusion-Driven Mechanisms of Protein Translocation on
Nucleic Acids .3. The Escherichia-Coli-Lac Repressor-Operator Interaction - Kinetic Measurements and
Conclusions. Biochemistry 20, 6961-6977, doi:10.1021/bi00527a030.
Zabet, N. R., 2012. System size reduction in stochastic simulations of the facilitated diffusion mechanism. Bmc
Systems Biology 6, doi:10.1186/1752-0509-6-121.
Zabet, N. R., Adryan, B., 2012. A comprehensive computational model of facilitated diffusion in prokaryotes.
Bioinformatics 28, 1517-1524, doi:10.1093/bioinformatics/bts178.
Zabet, N. R., Adryan, B., 2013. The effects of transcription factor competition on gene regulation. Frontiers in
genetics 4, 197.
Zabet, N. R., Adryan, B., 2015. Estimating binding properties of transcription factors from genome-wide binding
profiles. Nucleic Acids Research 43, 84-94, doi:10.1093/nar/gku1269.
Zabet, N. R., Foy, R., Adryan, B., 2013. The Influence of Transcription Factor Competition on the Relationship
between Occupancy and Affinity. Plos One 8, doi:10.1371/journal.pone.0073714.
27
|
1111.5814 | 1 | 1111 | 2011-11-24T16:44:22 | Stress Clamp Experiments on Multicellular Tumor Spheroids | [
"physics.bio-ph",
"physics.med-ph",
"q-bio.TO"
] | The precise role of the microenvironment on tumor growth is poorly understood. Whereas the tumor is in constant competition with the surrounding tissue, little is known about the mechanics of this interaction. Using a novel experimental procedure, we study quantitatively the effect of an applied mechanical stress on the long-term growth of a spheroid cell aggregate. We observe that a stress of 10 kPa is sufficient to drastically reduce growth by inhibition of cell proliferation mainly in the core of the spheroid. We compare the results to a simple numerical model developed to describe the role of mechanics in cancer progression. | physics.bio-ph | physics |
Stress clamp experiments on multicellular tumor spheroids
Fabien Montel1 , Morgan Delarue1 , Jens Elgeti1 , Laurent Malaquin1 , Markus Basan1 , Thomas Risler1 ,
Bernard Cabane3 , Danijela Vignjevic2 , Jacques Prost1,3 , Giovanni Cappello1 ,∗ and Jean-Fran¸cois Joanny1
1UMR 168, Institut Curie, Centre de Recherche, 26 rue d’Ulm 75005 Paris France.
2UMR 144, Institut Curie, Centre de Recherche, 26 rue d’Ulm 75005 Paris France. and
3ESPCI, 10 rue Vauquelin, 75005 Paris, France.
(Dated: December 20, 2013)
The precise role of the microenvironment on tumor growth is poorly understood. Whereas the
tumor is in constant competition with the surrounding tissue, little is known about the mechanics
of this interaction. Using a novel experimental procedure, we study quantitatively the effect of an
applied mechanical stress on the long-term growth of a spheroid cell aggregate. We observe that a
stress of 10kPa is sufficient to drastically reduce growth by inhibition of cell proliferation mainly in
the core of the spheroid. We compare the results to a simple numerical model developed to describe
the role of mechanics in cancer progression.
PACS numbers: 87.19.xj,87.19.R-,87.55.Gh
Cancer progression occurs in several stages.
In the
case of carcinomas, which are cancers of epithelial cells,
the primary tumor grows locally, until some cells invade
the neighboring tissue called the stroma which is essen-
tially made of extracellular matrix, fibroblast cells, im-
mune cells and capillary vessels. Three key elements con-
trol proliferation of the primary tumor: the accumulation
of gene mutations and the tumor biochemical and me-
chanical micro-environments. It is difficult to isolate in
vivo one of these factors to measure accurately its im-
portance. Several recent works suggest that mechanical
stress plays a role in tumor progression. A mechanical
stress applied to genetically predisposed tissues or tumor
spheroids grown in vitro induces signaling pathways that
are characteristic of cancer invasion [1, 2].
It has also
been shown that an increase of mechanical stress leads
to a reduction in cancer cell proliferation in vitro, and
drives apoptosis through the mitochondrial pathway [3–
5]. In spite of these experimental evidences, the precise
role of the micro-environment, and its interaction with
the tumor, are poorly understood. Our group has devel-
oped a theoretical framework [6–8] to describe the influ-
ence of the balance between cell division and apoptosis
on tumor growth under stress. The theory is based on
the existence of a homeostatic state of a tissue. This is
the steady state of the tissues where cell division bal-
ances cell death. The homeostatic stress is a function
of the biochemical state of the tissue and depends on
the local concentrations of nutrients, oxygen and growth
factors as well as on the environment of the tissue. Sig-
naling induced by the stroma can for example modify the
homeostatic state. In the simple case where the biochem-
ical state of the tissue can be maintained constant, the
homeostatic stress is the stress that the tissue can exert
at steady state on the walls of a confining chamber. It is
a measure of mechanical forces that cells can sustain in
this state. Indeed to grow against the surrounding tissue,
cells have to exert mechanical stress on the neighboring
cells.
In this paper we test experimentally the relevance of
the homeostatic stress concept. We measure the effect
of a known external stress on the growth of a cellular
aggregate mimicking a tumor over timescales longer than
the typical time scales of cell division or apoptosis. We
use a new experimental strategy to exert a well defined
mechanical stress on multicellular tumor spheroids for a
period of time exceeding 20 days.
We prepare colon carcinoma cell spheroids derived
from mouse CT26 cell lines (ATCC CRL-2638) using
a classical agarose cushion protocol [9]. The wells of a
48 wells plate are covered with agarose gel (Ultrapure
agarose, Invitrogen Co, Carlsbad, CA) and cell suspen-
sions are seeded on the gels at concentration of 20000 cells
per well. Cells self-assemble into spheroids in less than
24h. Cells are cultured under 95% air/ 5% CO2 atmo-
sphere in DMEM enriched with 10% calf serum (culture
medium). Using confocal microscopy, we check that the
shape of the spheroid is indeed close to a sphere. A con-
stant stress is applied on the tissue over long time scales
by imposing the osmotic pressure of a solution of the
bio-compatible polymer Dextran (Mw = 100kDa, Sigma-
Aldrich Co, St Louis, Mo). This polymer is known to be
neutral and is not metabolized by mammalians cells. We
also confirmed that it is neither a growth or a death factor
by plating cells for 3 days with Dextran and measuring
cell concentration and viability.
We first perform indirect stress measurements. A grow-
ing spheroid is positioned inside a closed dialysis bag (di-
ameter 10 mm, Sigma-Aldrich) which is then placed in
an external medium with added Dextran. The dialy-
sis membrane was chosen so that its molecular weight
cut-off (10 kDa) impedes the diffusion of Dextran. The
osmotic stress induces a force on the dialysis membrane,
which is transmitted in a quasi-static equilibrium to the
spheroid and calibrated as in [10, 11]. The stress ex-
erted on the cellular system can be seen as a network
stress that tends to reduce the volume occupied by the
spheroid. It acts directly on the cells, and not on the in-
terstitial fluid. The volume V (t)/V0 , normalized by the
initial volume of the spheroid V0 , is measured at succes-
sive times from a top view using differential interference
contrast microscopy (Axiovert 100, Zeiss). In the absence
of any applied stress, the spheroid reaches a steady state
with a typical diameter 900 µm. When Dextran is added
to the medium, a decrease of the growth rate dV
dt and of
the steady state volume are observed (Fig.1-Top). Inter-
estingly, after a stress release, the growth of the spheroid
resumes until it reaches the same steady state volume as
in the absence of external pressure. This indicates that
the effect of stress is fully reversible. Altogether these
results show that an external applied stress modulates
the growth of tumor spheroids.
We have also performed direct experiments where the
osmotic stress is applied onto the spheroid in the absence
of the dialysis membrane. In order to verify that Dextran
cannot diffuse inside the spheroid, we have placed it in
a medium supplemented with fluorescent FITC-Dextran
at an osmotic stress Π = 1000Pa. After 4 days of in-
cubation, the first 70 µm of the spheroid were imaged
using spinning disc microscopy. We measured that the
amount of Dextran able to penetrate into the spheroid is
negligible compared to the Dextran concentration in the
medium. The osmotic stress is thus applied on the first
layer of cells that plays the role of the dialysis membrane
in the direct experiment and transmits the stress to the
rest of the spheroid. The volume of the spheroid has also
been measured as a function of time (Fig.1-Bottom). We
observe a dependence of the growth rate and the steady
state size on stress very similar to that observed in the in-
direct experiment, validating our approach. Interestingly,
for a stress larger than 10kP a the effect of stress satu-
rates and the growth curves are indistinguishable from
each other.
The direct experiment is based on the application of a
mechanical stress on the surface of the spheroid through
an osmotic shock. Osmotic stress is known to have di-
rect effects on cell growth and apoptosis in particular
through the mitogen activated protein kinase (MAPK)
pathway [12–15]. However, in all these studies, the ef-
fect of an osmotic shock is only measured for an osmotic
stress two orders of magnitude larger than the one ap-
plied in our experiments (1M P a compared to 10kP a).
Moreover we do not observe any apoptosis at the sur-
face of the spheroid where the osmotic stress is exerted
(Fig. 2).
In addition, it can be checked by balancing
chemical potentials that the presence of Dextran outside
the spheroid creates a negligible concentration gradient
of all other solubles molecules. The concentration dif-
ference of a small soluble solute exchanged between the
interior and the exterior of the spheroid can be estimated
as ∆cs/cs = vΠ/kT (1 − Θs ) < 10−3 where cs and Θs are
the concentration and the volume fraction of the solute,
2
FIG. 1: Growth curves of individual spheroids under stress.
(Top) Normalized volume of individual spheroids as a func-
tion of time for the indirect experiments. The initial diameter
is Do = 350µm. At t = 12 days, stress is released. (Bot-
tom) Normalized volume of individual spheroids as a func-
tion of time for the direct experiments. The initial diameter
is Do = 200µm. The inserts show the principles of the two ex-
periments. The points are representative experimental data
for individual spheroids taken out of a larger set of experi-
ments. For each condition, N ≥ 3 experiments have been
recorded. The lines are the results of fits with the two rate
model. Error bars are the image analysis errors.
v the molecular volume of the solvent and Π the applied
osmotic stress. In other words, the chemical potential of
water in the cell is dominated by the small ions and it is
only slightly modified by the presence of Dextran.
Finally we investigate experimentally the spatial de-
pendence of cell division and apoptosis using cryo-
sections and immuno-fluorescence. Spheroids of com-
parable diameters are embedded in a freezing medium,
placed at −80oC and cut in slices of 5 µm thickness
at the level of their equatorial plane. Using a classi-
cal immuno-staining protocol, we then label fluorescently
dividing cells (in cyan with an anti-Ki67 antibody) and
apoptotic cells (in red with an anti -cleaved caspase3 an-
tibody) (See Fig. 2). We observe that in the absence
of external stress, cell division is distributed over all the
spheroid with an increase at the periphery, whereas for
an external stress of 1kPa, it is greatly reduced in the
center of the sections. As in previous studies [9, 16], we
observe an accumulation of apoptotic cells in the center
of the spheroid but with no measurable effect of stress on
this localization.
02468101214010203040506070Time (Day)Normalized Volume 0510152025300246810Normalized VolumeTime (day)pressure release0 Pa500 Pa0 Pa0 Pa0 Pa500 Pa2000 Pa5000 Pa10000 Pa20000 Pa3
FIG. 2: Effect of stress on the distribution of prolifera-
tion and apoptosis. Cryosections and immunofluorescence
of the spheroids are used to label the cell divisions (anti-
body against Ki67 in cyan) and apoptosis (antibody against
Cleaved-capase3 in red).(Left) Half section of a spheroid
grown in normal medium for 4 days (Right) Half section of a
spheroid grown with a stress of 1 kPa for 4 days
In order to better understand this stress dependence of
cell division and to interpret the generic trends of the ex-
perimental findings, we performed numerical simulations
similar to those of Ref.
[6, 8]. We adapt these simula-
tions to the geometry and setup of the experiments. In
brief, in the simulations, a cell is represented by a pair
of particles which repel each other and thus move apart.
When a critical distance is reached, the cell divides. After
division, each original particle constitutes, together with
a newly inserted particle in its surrounding, a daughter
cell. Particles belonging to different cells interact with
all particles with a short range interaction: a constant
attractive force describes cell-cell adhesion, while a repul-
sive short range potential ensures volume exclusion. The
viscous drag between cells is taken into account by a “Dis-
sipative Particle Dynamics“ -type thermostat. Finally a
constant apoptosis rate provides cell removal. In order
to mimic the experiments, tissue spheroids are grown in
a container together with a ”passive liquid” under stress.
This liquid interacts with the cell particles in a similar
way as cell particles with each others and it transmits
the stress on the spheroid.
As in the experiments, we observe a steady state that
depends on the applied stress. In the numerical simula-
tions each division event can be traced and the spatial
distribution of divisions can be measured to build virtual
cryosections of the simulated spheroids. We observe a
strong dependence of the division rate on the distance to
the surface of the spheroids and a clear decrease of cell
division anywhere in the section but stronger in the core
(see Fig.3).
Based on the growth curves and cryosection observa-
tions, we present a simple two rate description of the
FIG. 3: Dissipative Particle Dynamics Simulations.
(Left)
Bulk division rate kd , surface rate increment δks and apop-
tosis rates ka as function of the applied pressure P . (Right)
Virtual cryosections of the simulated spheroids for an external
pressure P=0 Pa or P=5kPa. The pressure units have been
calibrated from the experiments using the saturation pressure.
spheroid growth in the absence or the presence of exter-
nal stress: the core of the spheroid is mostly undergoing
apoptosis whereas its periphery is proliferating. In this
situation, the net growth rate is proportional to the area
(∝ r2 ) while the net death rate is proportional to the vol-
ume (∝ r3 ). This surface growth effect leads to a stable
steady state size. The surface localization of the prolif-
eration can be obtained using purely mechanical consid-
erations. A cell must deform its environment to grow.
The deformation is facilitated if the cell is closer to the
surface, and this implies that proliferation is favored at
the surface. The increased number of cell divisions at
the surface drives a flow from the surface of the spheroid
toward its center. The flow is a possible explanation for
the accumulation of the long lasting apoptotic markers
in the center of the spheroid. A mechanical control of
cell cycle entrance can also explain the growth of tumor
spheroids in free suspension [17].
In this case nutrient
depletion causes the formation of a necrotic core which
generates death at the inner surface of the viable rim
(∝ r2 ) i.e. not proportional to the volume thereby not
generating a steady state [17, 18]. Using a fluorescently
labelled growth factor (Alexa 555- EGF) we have verified
that the transport of these molecules is not affected by
stress. This result supports a direct mechanical effect on
the division rate.
Our two rate model can be seen as a simplified version
of the two rate model of Radszuweit et al. [19]. The net
bulk growth rate is k = kd − ka , where kd and ka are the
division and apoptosis rates respectively. It is a function
of stress. At the surface, the net growth rate kd − ka +
δks is larger and δks has a different stress dependence.
Taking into account surface and bulk growth, the growth
equation reads :
∂tN = (kd − ka )N + δksNs
Assuming a constant cell density and a constant thickness
λ of the region where the division rate increment is equal
(1)
Ki-67cleaved Caspase-3 P = 0 PaP = 1 kPa100µm 0.01 0.1 1 10 0 2 4 6 8 10 12 14 16k / kaP (kPa)δkskdkaP= 0 PaP= 5 kParecently dividedcellnormalcell4
mechanical effects can have strong implications in cancer
proliferation. This raises the question of the players in
the crosstalk between stress and cellular response and in
particular of the nature of the stress sensor.
We would like to thank F. Brochard, F. Graner, P.
Nassoy, K. Alessandri and J. Kaes for useful discussions.
F.M. and G.C. would like to thank Axa Research Fund
and CNRS for funding. The group belongs to the CNRS
consortium CellTiss.
∗ Electronic address: [email protected]
[1] J. Whitehead, D. Vignjevic, C. Futterer, E. Beaurepaire,
S. Robine, and E. Farge, HFSP journal 2, 286 (2008),
ISSN 1955-2068.
[2] Z. Demou, Annals of Biomedical Engineering 38, 3509
(2010), ISSN 0090-6964.
[3] G. Cheng, J. Tse, R. K. Jain, and L. L. Munn, PloS one
4, e4632 (2009), ISSN 1932-6203.
[4] T. Roose, P. Netti, L. Munn, Y. Boucher, and R. Jain,
Microvascular research 66, 204 (2003), ISSN 0026-2862.
[5] G. Helmlinger, P. Netti, H. Lichtenbeld, R. Melder, and
R. Jain, Nature Biotechnology 15, 778 (1997), ISSN
1087-0156.
[6] M. Basan, T. Risler, J.-F. Joanny, X. Sastre-Garau, and
J. Prost, HFSP journal 3, 265 (2009), ISSN 1955-205X.
[7] J. Ranft, M. Basan, J. Elgeti, J.-F. Joanny, J. Prost,
and F. Julicher, Proceedings of the National Academy
of Sciences of the United States of America 107 (2010),
ISSN 1091-6490.
[8] M. Basan, J. Prost, J.-F. Joanny, and J. Elgeti, Physical
biology 8, 026014 (2011), ISSN 1478-3975.
[9] W. Mueller-klieser and L. A. Kunz-schughart, Journal of
Biotechnology (2010), ISSN 0168-1656.
[10] C. Bonnet-Gonnet, L. Belloni, and B. Cabane, Langmuir
10, 4012 (1994), ISSN 0743-7463.
[11] A. Bouchoux, P.-E. Cayemitte, J. Jardin, G. G´esan-
Guiziou, and B. Cabane, Biophysical journal 96, 693
(2009), ISSN 1542-0086.
[12] K. J. Cowan, Journal of Experimental Biology 206, 1107
(2003), ISSN 00220949.
[13] B. Racz, D. Reglodi, B. Fodor, B. Gasz, A. Lubics,
F. Gallyas, E. Roth, and B. Borsiczky, Bone 40, 1536
(2007), ISSN 8756-3282.
[14] M.-B. Nielsen, S. T. Christensen, and E. K. Hoffmann,
American journal of physiology. Cell physiology 294,
C1046 (2008), ISSN 0363-6143.
[15] Y. Xie, W. Zhong, Y. Wang, A. Trostinskaia, F. Wang,
E. E. Puscheck, and D. A. Rappolee, Molecular human
reproduction 13, 473 (2007), ISSN 1360-9947.
[16] W. Mueller-Klieser, American Journal of Physiology-
Cell Physiology 273, C1109 (1997), ISSN 0363-6143.
[17] D. Drasdo and S. Hohme, Physical biology 2, 133 (2005),
ISSN 1478-3975, URL http://www.ncbi.nlm.nih.gov/
pubmed/16224119.
[18] G. Schaller and M. Meyer-Hermann, Physical Review E
71, 1 (2005), ISSN 1539-3755, URL http://link.aps.
org/doi/10.1103/PhysRevE.71.051910.
[19] M. Radszuweit, M. Block, J. Hengstler, E. Scholl, and
D. Drasdo, Physical Review E 79, 051907 (2009).
FIG. 4: Evolution of growth rates with stress. (Left) Surface
division rate increment δks as a function of stress. (Right)
Bulk growth rate k as a function of stress. For each condi-
tion, N ≥ 3 experiments have been recorded. The errors bar
are obtained using a jackknifing method and represent the
efficiency of the fitting algorithm
to δks , one can express for R > λ the rate of volume
increase as :
∂tV = (kd − ka )V + (36π)1/3 δksλV 2/3
(2)
For small spheroids (R << λ) the growth rate is posi-
tive and constant (kd + δkd − ka ), this leads to the previ-
ously described exponential growth [17, 18]. In our case
the growth curves can readily be fitted by Eq. 2 in the
range of large spheroids [20]. The variation with pressure
of the parameters k = kd − ka and δks is given in Fig. 4.
The surface growth rate δks is less affected by stress than
the bulk growth rate k . A similar fit can be performed on
the simulations. The bulk and surface growth rates k and
δks are represented on Fig. 3. Although both the surface
and bulk growth rates depend exponentially on pressure,
the decay constant of the surface rate is much smaller
than that of the bulk rate. While the bulk rate decreases
by more than one order of magnitude, the surface rate
decreases by a factor 3. This supports the hypotheses
of the simulations. In summary, cell division in the core
of the spheroid is strongly affected by stress whereas cell
division rate increment on the surface of the spheroid
depends more weakly on stress.
In conclusion, we have shown by a direct measure-
ment of the tissue response to an external stress that
the application of an external stress drastically limits
the growth of tumoral spheroids. Previous approaches
[5, 21] had used the elastic deformation of poro-elastic
gels to measure the maximum stress that can be devel-
oped by spheroids. The measured stress in these experi-
ments is in the same range as in our measurements. In a
recent study a localized increase of mitochondrial apop-
tosis and a reduction of proliferation in presence of stress
were also reported [3]. This difference with our results
may be due to the fact that in our case we are not con-
trolling the rigidity of the surrounding substrate but the
applied pressure. This may leads to a different response
of the spheroid. Our results favors the idea that direct
0500010000-0.55-0.5-0.45-0.4-0.35-0.3-0.25-0.2Pressure (Pa)k = kd - ka05000100000.20.250.30.350.40.450.50.55Pressure (Pa)δks[20] L. Von Bertalanffy, The Quarterly Review of Biology
32, 217 (1957), URL http://www.jstor.org/stable/
2815257.
[21] A. Fritsch, M. Hockel, T. Kiessling, K. D. Nnetu, F. Wet-
zel, M. Zink, and J. A. Kas, Nat Phys 6, 730 (2010).
5
|
1004.0973 | 1 | 1004 | 2010-04-06T21:00:57 | Response of a Hodgkin-Huxley neuron to a high-frequency input | [
"physics.bio-ph",
"q-bio.NC"
] | We study the response of a Hodgkin-Huxley neuron stimulated by a periodic sequence of conductance pulses arriving through the synapse in the high frequency regime. In addition to the usual excitation threshold there is a smooth crossover from the firing to the silent regime for increasing pulse amplitude $g_{syn}$. The amplitude of the voltage spikes decreases approximately linearly with $g_{syn}$. In some regions of parameter space the response is irregular, probably chaotic. In the chaotic regime between the mode-locked regions 3:1 and 2:1 near the lower excitation threshold the output interspike interval histogram (ISIH) undergoes a sharp transition. If the driving period is below the critical value, $T_i < T^*$, the output histogram contains only odd multiples of $T_i$. For $T_i > T^*$ even multiples of $T_i$ also appear in the histogram, starting from the largest values. Near $T^*$ the ISIH scales logarithmically on both sides of the transition. The coefficient of variation of ISIH has a cusp singularity at $T^*$. The average response period has a maximum slightly above $T^*$. Near the excitation threshold in the chaotic regime the average firing rate rises sublinearly from frequencies of order 1 Hz. | physics.bio-ph | physics |
Response of a Hodgkin-Huxley neuron to a high-frequency input
Faculty of Physics, Adam Mickiewicz University, Umultowska 85, 61-614 Poznan, Poland
L. S. Borkowski
We study the response of a Hodgkin-Huxley neuron stimulated by a periodic sequence of conduc-
tance pulses arriving through the synapse in the high frequency regime. In addition to the usual
excitation threshold there is a smooth crossover from the firing to the silent regime for increas-
ing pulse amplitude gsyn. The amplitude of the voltage spikes decreases approximately linearly
with gsyn.
In some regions of parameter space the response is irregular, probably chaotic. In the chaotic
regime between the mode-locked regions 3:1 and 2:1 near the lower excitation threshold the output
interspike interval histogram (ISIH) undergoes a sharp transition. If the driving period is below
the critical value, Ti < T ∗, the output histogram contains only odd multiples of Ti. For Ti > T ∗
even multiples of Ti also appear in the histogram, starting from the largest values. Near T ∗ the
ISIH scales logarithmically on both sides of the transition. The coefficient of variation of ISIH has
a cusp singularity at T ∗. The average response period has a maximum slightly above T ∗. Near the
excitation threshold in the chaotic regime the average firing rate rises sublinearly from frequencies
of order 1 Hz.
I.
INTRODUCTION
Biological neurons transmit information in the form of
sharp spikes of potential difference across the lipid bi-
layer forming the wall of the nerve cell. This feature of
the cell's reaction to input signals is remarkably consis-
tent in different organisms and different types of neurons.
The action potential spikes are assumed to be the prin-
cipal carrier of information. The early view that infor-
mation is transmitted via rate coding has evolved. It is
now recognized that also the spike time coding is used
in neural systems1,2. While the precise coding recipe is
unknown it is clear that the knowledge of the response
of various types of neurons to different stimuli is funda-
mental to formulating the theory of information transfer
in the neural system.
Our understanding of conductance-based models of
neurons is largely based on the Hodgkin-Huxley (HH)
model originally formulated to describe the dynamics of
the membrane potential of the squid giant axon3. The de-
tailed voltage-clamp measurements of the voltage-gated
potassium and sodium ion currents led to revisions of the
HH model. The modifications required to achieve bet-
ter agreement with experiments were reviewed by Clay4.
Studies of single neurons and neuronal networks often
employ simplified models, such as integrate-and-fire and
FitzHugh-Nagumo (FHN) models5,6. It is believed that
the two-dimensional flow models such as FHN reproduce
qualitatively the behavior of the HH model. However
these simplifications are not always justifiable7 -- 9. In an
interesting analysis of chaos in the HH model Gucken-
heimer and Oliva9 point out that even the concept of a
firing threshold may be more subtle that just a smooth
hypersurface dividing subthreshold and suprathreshold
membrane potentials.
Over the years many studies of HH equations were
carried out,
including stochastic variations of various
quantities10 -- 12. An important question is to what extent
the qualitative properties of neuron response depend on
the functional form of the input signal. One frequently
used form of input is constant plus a sinusoidal term.
However the physiological signals are more pulse-like. In
a strongly nonlinear system this may lead to substantial
differences in the output.
In the sinusoidally driven HH model the excitation
threshold rises sharply at large frequencies. The phase
diagram in the frequency-current amplitude plane con-
sists of three phase locked regions with integer ratio of
the output period to the input period, ¯To/Ti, 1:1, 2:1,
and 3:1. There are also areas of fractional locking and
bistable or chaotic response around these phase-locked
states13 -- 15.
It was pointed out that the edges of mode-locked
plateaus have analogies to phase transitions in the equi-
librium statistical mechanics. Two forms of scaling of the
average deviation from perfect mode-locking were found
near the edges of plateaus with constant p/q, where p
and q are integers, indicating number of input spikes per
number of output action potentials16. The scaling has
either exponent 1/2 or is logarithmic. In this paper we
will show that scaling is more common and appears also
near the multimodal transition points.
Here we assume the α form of postsynaptic current,
Isyn ∼ t exp(−t/τ ), where t is time from the onset of
the input spike and τ is the time scale of the synaptic
action. This form is close to experimental observation
although it does not take into account a more complex
dynamics of the ion channel kinetics, usually described
in the Markovian scheme.
The general form of the phase diagram of the Hodgkin-
Huxley model with this input was studied initially in
Ref.17. However many important questions are still to
be answered. One of them is the behavior of the system
in the high-frequency limit. In the following we present
the model and show the main features of high-frequency
response.
2
II. THE MODEL
The Hodgkin-Huxley neuron subject to periodic con-
ductance pulses is defined by the following set of
equations,3
CdV /dt = −gN am3h(V − VN a) − gKn4(V − VK)
−gL(V − VL) + Iext + Isyn,
dm/dt = −(am + bm)m + am,
dh/dt = −(ah + bh)h + ah,
dn/dt = −(an + bn)n + an,
(1)
(2)
(3)
(4)
where
FIG. 1: Sample voltage trace for a constant input current
Iext = 10µA/cm2.
am = 0.1(V + 40)/[1 − e−(V +40)/10],
(5)
bm = 4e−(V +65)/18,
ah = 0.07e−(V +65)/20,
bh = 1/[1 + e−(V +35)/10],
(6)
(7)
(8)
an = 0.01(V + 55)/[1 − e−(V +55)/10],
(9)
bn = 0.125e−(V +65)/80.
(10)
In equations (5)-(10) the voltage is expressed in mV
and the rate constants α and β are given in ms−1.
The reversal potentials of sodium, potasium and leak-
age channels are VN a = 50mV, VK = −77mV, and
VL = −54.5mV, respectively. The corresponding max-
imum conductances are gN a = 50mV, gK = 36mS/cm2,
and gL = 0.3mS/cm2. The capacity of the membrane is
C = 1µF/cm23.
The synaptic current Isyn is given by the following
equation,
Isyn(t) = gsyn X
n
α(t − tin)(Va − Vsyn),
(11)
FIG. 2: The ratio of the average output spiking rate to the
input rate, k = ¯To/Ti. Mode-locked regions with k = 1, 2, 3
and k = 4, 5, 6, 7, 8 are shown in black and grey respectively.
Voltage peaks were counted as spikes when V exceeded 0. For
high values of gsyn the neuron does not respond.
where tin denotes the start of the nth pulse, gsyn is the
conductivity of the synapse, Va = 30mV is the maximum
potential in the postsynaptic area and Vsyn = −50mV is
the reversal potential of the synapse. The period of the
synaptic drive is Ti = tin+1 − tin. The external current
Iext is set to 0, except for a sample run shown in Fig. 1.
The time-dependence is given by the function
α(t) = (t/τ )e−t/τ )Θ(t),
(12)
where τ is time scale characterizing the dynamics of the
synaptic action and Θ(t) is the Heaviside step function.
We study the dependence of the output interspike sepa-
ration To on Ti and gsyn.
3
FIG. 3: Values of maxima and minima of the membrane potential V (t) as a function of synaptic conductivity gsyn for input
spike intervals Ti = 2.5ms, 3.5ms, and 4.5ms.
Equations (1)-(10) were integrated with the fourth or-
der Runge-Kutta scheme. The time step was 0.01 ms.
For each parameter set the simulation was run for 30
seconds. Results of the initial three seconds of each data
set were discarded to avoid transient behavior.
In the
chaotic regime the data were obtained from five runs for
each value of the horizontal coordinate.
III. RESULTS
The average output spiking rate in the form of a color
map as a function of the input period Ti and max-
imum synaptic conductivity gsyn is presented in Fig.
2. The mode-locked regions are shown as areas of uni-
form color. For small Ti the total incoming current is
approximately constant with a small modulation, and
the excitation threshold rises linearly with increasing Ti,
gsyn ≃ 0.04Ti mS/(ms cm2). For gsyn exceeding approx-
imately 0.4Ti mS/(ms cm2) the spiking action does not
occur. We can see from Fig. 2 that this behavior sets in
below Ti ≃ 6ms .
The obtained phase diagram is qualitatively different
from a response to a sinusoidal input, where the excita-
tion threshold diverges as 1/Ti, for Ti → 0. In general
we may expect that the constraint of charge balancing,
R t+Ti
Idt = 0, will have a significant impact at high fre-
t
quencies. For intermediate values of the input period,
5ms < Ti < 13ms, the topology of the phase diagram
resembles results obtained with sinusoidal input, see e.g.
Fig. 2 of Ref.13.
FIG. 4: For high synaptic conductivities the distinction be-
tween action potential and the background oscillations loses
its meaning. This sample was obtained for Ti = 4.5ms and
gsyn = 2.35mS/cm2.
Fig. 3 shows dependence of minima and maxima of
V on gsyn for three input frequencies. The amplitude of
response decreases linearly with increasing gsyn. There is
no well-defined spiking threshold. There are intervals of
parameter values for which the response is highly irregu-
lar and the values of maxima and minima of V vary sig-
nificantly.
4
(a)The spectrum of interspike separations of the
FIG. 5:
output signal as a function of the input period Ti for gsyn =
0.4mS/cm2, (b) Detailed view of the chaotic region between
Ti = 5ms and 6ms. Each ISI cluster belongs to different k,
where k = 2, 3, 4, 5, .... The distinction between k = 2 and
k = 3 is blurred.
FIG. 6: Scaling of the excitation edge of (a) odd-only mul-
tiples of the input Ti, and (b) all integer multiples, in the
chaotic region between k = 2 and k = 3. For g = 0.2mS/cm2,
the transition occurs at T = 6.54175ms.
A sample time-dependence of the membrane potential
is shown in Fig. 4. The maxima of V span almost the
entire range between −60mV and 0mV. There is no clear
separation of spikes from the rest of the signal.
Chaotic behavior in the parameter space between the
3:1 and 2:1 mode-locked regions leads to multimodal re-
sponse. The interspike separation for gsyn = 0.4mS/cm2
is shown in Fig. 5. For Ti between 5.5ms and 6ms all
integer multiples of input Ti with the exception of the
lowest one appear in the output ISIH.
It is interesting to note that ISI histograms (ISIH) from
some older experiments on nerve fibers of monkeys18 and
single neurons in the primary visual cortex of a cat19
show some similarity to Fig. 5. Experimental histograms
are sequences of diminishing peaks occuring at integer
multiples of the input interspike separation. In Fig. 5
the lowest element of the sequence is missing due to the
refractoriness of the neuron. Similar form of ISIH was ob-
tained in a theoretical study of a bistable system stimu-
lated by periodic function with additive Gaussian noise,20
where the presence of noise was essential. However the
multimodal histogram was also obtained in a simulation
of a deterministic modification of the HH model21.
The HH model studied here does not contain stochastic
terms. The multimodal response in Fig. 5 is a result of
a deterministic nonlinearity. Thus noise is not the only
ingredient enabling the reproduction of the multimodal
experimental ISIH. It is possible to identify the source of
multimodality by studying ISIH in more detail.
5
FIG. 7: The multimodal transition at g = 0.17mS/cm2.
FIG. 9: The ratio k = To/Ti for g = 0.17mS/cm2. The
maximum of k is shifted approximately 0.2 ms to the right
relative to maximum of CV (see Fig. 8).
FIG. 8: Coefficient of variation for g = 0.17mS/cm2. The
variability near Ti = 6.6ms is due to the proximity to the
firing threshold.
FIG. 10: The location of the multimodal transition (filled
squares) on the response diagram.
Close
to the
excitation threshold,
at gsyn ≃
0.2mS/cm2, there exists a transition from the odd-only
ISIH to ISIH with all integer multiples of Ti, see Fig. 6.
Near the transition the edges of high-k clusters scale loga-
rithmically. The scaling holds both along the Ti axis and
along the gsyn axis. It can be viewed as a competition
between the odd and the even multiples of the driving
period.
A clear indication of this "spectral" transition is the
singular behavior of the coefficient of variation, see Fig.
8. At the transition CV is of order 1. and k is signif-
icantly larger than 3. The maximum k occurs approx-
imately 0.2 ms above the singularity of CV. One may
also think of this shift as a result of relaxation from the
constraint of odd-only modes below T ∗. At T ∗ the high-
est even modes become available and this leads to the
increase of k.
If such transition were found experimentally it would
be a clear sign of the deterministic nonlinear dynam-
ics.
In the presence of noise this sharp feature would
be smeared and would vanish if noise dominates the dy-
namics of the system.
6
frequency) bands vanish logarithmically near the line of
critical points (gsyn,T ∗). The firing rate has a minimum
at Ti ≃ T ∗ + 0.2ms. Periodically stimulated giant axons
of squid have similar nonmonotonic dependence of the
firing rate on the current pulse amplitude between the
k = 2 and k = 3 states24. This experiment also showed
linear dependence of the firing rate on pulse amplitude
near the threshold for Ti > T ∗, similarly to Fig. 11.
Although the experimental pulses were rectangular, dif-
ferent from the α(t) form with an exponential tail, the
qualitative features do not depend much on the precise
shape of a pulse. For short pulses the neuron's reaction is
determined mainly by the time integral of the stimulus.
The multimodal response occuring in certain sen-
sory neurons may result from noise20 or deterministic
nonlinearity25. It would be interesting to look for exper-
imental evidence of the odd-all transition.
It found, it
would be a clear evidence that the neuron dynamics is
dominated by nonlinearity, not noise.
The behavior of the model at small Ti may be useful to
both coincidence detection and estimation of the signal
strength. The optimal sensitivity in this case is inversely
proportional to frequency.
Our calculation also supports the view expressed by
authors of Ref.9 that boundaries between various parts
of the response diagram are not always clear-cut and may
form complicated patterns. This statement also applies
to the excitation threshold in the chaotic regime.
In the Hodgkin's classification of intrinsic excitability26
class 1 neurons maintain firing at arbitrarily low fre-
quencies in response to weak inputs and have continu-
ous frequency-current (f -I) curve. Class 2 neurons fire
with certain relatively large frequency, usually of order
40-50 Hz, when stimulus exceeds threshold and have a
discontinuous f -I curve. Class 1 and class 2 neurons
sometimes are described as integrators and resonators
respectively27. According to the commonly held view
a neuron cannot be an integrator and resonator at the
same time. However we showed that the deterministic
HH neuron in a chaotic regime near excitation thresh-
old may oscillate with arbitrarily small frequencies and
may perform integration at time scales much longer than
the period of its main resonance. The character of the
response depends strongly on the functional form of the
stimulus and parameters of the model. A recent study
showed that the same pyramidal neurons behave as inte-
grators in vitro and resonators in vivo.28
The multimodal response of the HH neuron near 140-
180 Hz is not a typical resonance since no particular fre-
quency is preferred. The multiples of the driving fre-
quency alternate chaotically. The average output fre-
quency depends nonmonotonically on the stimulus am-
plitude. Similar nonmonotonic f vs. I relation was found
in periodically stimulated giant axons of squid24. Smaller
stimuli favor higher multiples of the driving period. Stud-
ies of large neuronal networks of various types suggest
that there may be a complex interplay between the inte-
grating behavior and the resonant action.22
synaptic conductivity for Ti =
FIG. 11: Frequency vs.
7.45ms. Each data point is averaged over 15 runs for 60 s
with different initial conditions. The initial 6 s from each run
were discarded.
IV. CONCLUSIONS
For high synaptic drives in the high frequency regime
distinguishing the action potential from the background
activity becomes problematic. In this limit the neuron is
very sensitive to small changes of the functional form of
the signal. For periodic drive with small time constantτ
and Ti below 6ms the width of the spiking regime along
the gsyn axis scales linearly with Ti. The quality of the
neuron's response deteriorates linearly with increasing
gsyn. This is in contrast to findings for a sinusoidal sig-
nal, and more generally for a class of signals satisfying the
constraint of charge balancing, where the spiking action
remains well defined in the high-frequency limit.
A mechanism of suppression of the neuron's activity
might help explain self-regulating behavior of neocorti-
cal networks. Various mechanisms of homeostatic action
for neural microcircuits were proposed.22 It would be use-
ful to investigate whether more realistic extensions of the
Hodgkin-Huxley model also exhibit self-regulation in re-
sponse to high-frequency inputs. The network of such
neurons would have a "safety switch" built in at the level
of individual cells. For Ti between 4 and 6 ms the up-
per critical synaptic conductivity is of order 2mS/cm2,
which is in the realistic range for neocortical pyramidal
neurons23.
The input ISI of 4 − 8ms is important to understand-
ing the dynamics of the Hodgkin-Huxley model. In the
chaotic region between the k = 2 and k = 3 locked states
the coefficient of variation of ISI has a singularity at the
transition between the odd-only and all-integer multi-
ples of the driving period. The odd modes dominate
in the vicinity of the k = 3 state. The low-k (high-
The ability to precisely control the nerve cell's po-
tential oscillations is important in constructing de-
vices performing the procedure known as Deep Brain
Stimulation29 -- 31, which operate at frequencies above
100 Hz. While our model does not satisfy the charge-
balancing constraint required in the stimulation of in-
vivo systems, we believe the present study improves our
understanding of high-frequency neural oscillators.
Acknowledgments
7
The author thanks J. W. Mozrzymas, D. W´ojcik, K.
Bodova, T. Burwick, and P. Suffczy´nski for discussions.
Computations were performed in the Computer Center
of the Tri-city Academic Computer Network in Gdansk.
1 T. J. Sejnowski, Nature (London) 376, 21 (1995).
2 D. Ferster and N. Spruston, Science 270, 756 (1995).
3 A. L. Hodgkin and A. F. Huxley, J. Physiol. (London) 117,
500 (1952).
4 J. R. Clay, Prog. Biophys. Mol. Biol. 88, 59 (2005).
5 W. Gerstner, Phys. Rev. E 51, 738 (1995).
6 P. C. Bressloff and S. Coombes, Phys. Rev. Lett. 81, 2168
(1998).
7 J. Rinzel and R. Miller, Math. Biosci. 49, 27 (1980).
8 D. Brown, J. Feng, and S. Freerick, Phys. Rev. Lett. 82,
4731 (1999).
17 H. Hasegawa, Phys. Rev. E 61, 718 (2000).
18 J. E. Rose, J. F. Brugge, D. J. Anderson, and J. E. Hind,
J. Neurophysiol. 30, 769 (1967).
19 R. M. Siegel, Physica 42D, 385 (1990).
20 A. Longtin, A. Bulsara, and F. Moss, Phys. Rev. Lett. 67,
656 (1991).
21 J. R. Clay, J. Comput. Neurosci. 15, 43 (2003).
22 R. C. Muresan and C. Savin, J. Neurophysiol. 97, 1911
(2007).
23 N. Ho and A. Destexhe, J. Neurophysiol. 84, 1488 (2000).
24 N. Takahashi, Y. Hanyu, T. Musha, R. Kubo, and G. Mat-
9 J. Guckenheimer and R. O. Oliva, SIAM J. Appl. Dyn.
sumoto, Physica D 43, 318 (1990).
Sys. 1, 105 (2002).
10 S. G. Lee and S. Kim, Phys. Rev. E 60, 826 (1999).
11 E. V. Pankratova, A. V. Polovinkin, and E. Mosekilde,
Eur. Phys. J. B 45, 391 (2005).
12 S. Luccioli, T. Kreuz, and A. Torcini, Phys. Rev. E 73,
041902 (2006).
13 S. G. Lee and S. Kim, Phys. Rev. E 73, 041924 (2006).
14 D. T. W. Chik, Y. Wang, and Z. D. Wang, Phys. Rev. E
64, 021913 (2001).
15 Y.-Q. Che, J. Wand, W.-J. Si, and X.-Y. Fei, Chaos, Soli-
tons and Fractals 39, 454 (2009).
25 D. T. Kaplan, J. R. Clay, T. Manning, L. Glass, M. R.
Guevara, and A. Shrier, Phys. Rev. Lett. 76, 4074 (1996).
26 A. L. Hodgkin, J. Physiol. 107, 165 (1948).
27 E. M. Izhikevich, Int. J. Bif. Chaos 10, 1171 (2000).
28 S. A. Prescott, S. Ratt´e, Y. D. Koninck, and T. J. Se-
jnowski, J. Neurophysiol. 100, 3030 (2008).
29 A. L. Benabid, P. Pollak, C. Gervason, D. Hoffmann, D. M.
Gao, M. Hommel, J. E. Perret, and J. de Rougemont,
Lancet 337, 403 (1991).
30 R. E. Gross and A. M. Lozano, Neurol. Res. 22, 247 (2000).
31 C. C. McIntyre, M. Savasta, B. L. Walter, and J. L. Vitek,
16 J. R. Engelbrecht and R. Mirollo, Phys. Rev. E 79, 021904
J. Clin. Neurophysiol. 21, 1 (2004).
(2009).
|
1711.02175 | 1 | 1711 | 2017-11-02T14:43:20 | Determining the Refractive Index of Human Hemoglobin Solutions by Kramers-Kronig Relations with an Improved Absorption Model | [
"physics.bio-ph",
"physics.med-ph",
"physics.optics"
] | The real part of the refractive index (RI) of aqueous solutions of human hemoglobin is computed from their absorption spectra in the wavelength range $250\,{\rm nm} - 1100\,{\rm nm}$ using the Kramers-Kronig (KK) relations and the corresponding uncertainty analysis is provided. The strong ultraviolet (UV) and infrared absorbance of the water outside this spectral range were taken into account in a previous study employing KK relations. We improve these results by including the concentration dependence of the water absorbance as well as by modeling the deep UV absorbance of hemoglobin's peptide backbone. The two free parameters of the model for the deep UV absorbance are fixed by a global fit. | physics.bio-ph | physics | Research Article
Applied Optics
1
7
1
0
2
v
o
N
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
5
7
1
2
0
.
1
1
7
1
:
v
i
X
r
a
Determining the Refractive Index of Human
Hemoglobin Solutions by Kramers-Kronig Relations
with an Improved Absorption Model
JONAS GIENGER1,*, HERMANN GROSS1, JÖRG NEUKAMMER1, AND MARKUS BÄR1
1Physikalisch-Technische Bundesanstalt (PTB), Abbestrasse 2–12, 10587 Berlin, Germany
*Corresponding author: [email protected]
Compiled November 8, 2017
The real part of the refractive index (RI) of aqueous solutions of human hemoglobin is computed from
their absorption spectra in the wavelength range 250 nm–1100 nm using the Kramers-Kronig (KK) rela-
tions and the corresponding uncertainty analysis is provided. The strong ultraviolet (UV) and infrared
absorbance of the water outside this spectral range were taken into account in a previous study employ-
ing KK relations. We improve these results by including the concentration dependence of the water
absorbance as well as by modeling the deep UV absorbance of hemoglobin's peptide backbone. The two
free parameters of the model for the deep UV absorbance are fixed by a global fit.
© 2016 Optical Society of America. One print or electronic copy may be made for personal use only. Systematic reproduction and distribution, duplication
of any material in this paper for a fee or for commercial purposes, or modifications of the content of this paper are prohibited.
OCIS codes:
Dispersion; 300.1030 Absorption
000.1430 Biology and medicine; 000.3860 Mathematical methods in physics; 120.4530 Optical constants; 260.2030
http://dx.doi.org/10.1364/AO.55.008951
1. INTRODUCTION
The optical properties of biological cells and tissues have been
subject to research for many decades. Refractometry in cells is
used, e. g., for protein detection [1]. Recently, the refractive in-
dex (RI) or RI distribution of cells has been measured by phase
microscopy [2], holographic techniques [3], absorption cytome-
try [4] or optical tomography [5]. From such measurements, the
dry mass or concentrations of proteins in the cell can be derived,
provided the relation to the optical properties is known.
Analysis of blood samples includes the determination of the
quantities of the so called complete blood count (CBC), one
of the most frequently performed measurements in laboratory
medicine. Besides the concentrations of red blood cells (RBCs,
erythrocytes), white blood cells and platelets, an important in-
dicator for diseases, e. g., anemia, is the mean cellular volume
(MCV) of red blood cells. Besides the calculation of the MCV
as ratio of the hematocrit value and the RBC concentration,
flow-cytometric detection of light scattering by RBCs is being
used for more than three decades to determine the volume
and hemoglobin content of individual cells [6, 7] at a through-
put in the range of 1000 events per second. Of course, a de-
tailed knowledge of the cells' refractive index is required for
this. Since RBCs are mainly composed of the protein complex
hemoglobin (Hb), which is dissolved in water at an average in-
tracellular concentration of typically 340 g L−1 for healthy per-
sons, the knowledge of the Hb RI is of crucial importance to
reliably calculate RBC volume and Hb content.
Within the blood of a single person, the intracellular Hb con-
centration fluctuates among the individual RBCs, with a coeffi-
cient of variation around 6–8% [5, 6]. Because of the significant
influence of this variation, it does not suffice to know the RI of
hemoglobin solutions at the mean concentration, but the depen-
dence of the RI on the concentration has to be specified.
The absorption spectra of Hb solutions were measured with
high accuracy over a wide range of wavelengths and are known
for several decades. In contrast, measurements of their refrac-
tive index, especially at physiologically relevant high concen-
trations, are challenging and have only been presented as late
as 2005 for a wide spectral range [8]. However, these data have
much larger measurement uncertainties than the correspond-
ing absorption spectra, such that the real part of the complex re-
fractive index is less accurately known than its imaginary part,
the absorbance.
As a remedy several authors [9, 10] suggested to employ
Kramers-Kronig (KK) relations in order to obtain the real part of
the RI directly from accurate measurements of absorption spec-
tra, i. e., the imaginary part of the RI as a function of wavelength.
However, this is a non-trivial task due to the finite wavelength
range of the measured spectra and the global, long-ranged char-
Research Article
acter of the KK relations which yield the real part of the RI as
an integral transformation of the imaginary part. In this paper,
we supplement and extend the literature spectra [8] with an ab-
sorption model for proteins towards the UV. As a result we ob-
tain RIs that are in better agreement with results obtained by re-
flectance measurements for oxyhemoglobin solutions [11]. Fur-
thermore we obtain results for the RI of deoxyhemoglobin and
give the refractive increment, i. e., the slope of the RI with respect
to concentration for the first time. We also extend the previ-
ous treatment by considering the measurement uncertainties of
the obtained RIs. These uncertainties result from the propaga-
tion of the uncertainties of the measured absorption spectra [8]
and the refractive index data obtained by reflectance measure-
ments [11], to which we fit the two free parameters of the deep
UV absorption model. The mathematical model for the absorp-
tion spectra and related method are presented in sections 2 and
3. Our approach can be applied to any variant of hemoglobin.
Section 2 A describes the imaginary part of the RI of a Hb solu-
tion in dependence on concentration. KK relations are used in
section 2 B to obtain an expression for the real part of the solu-
tion's RI. A model for the deep UV absorbance of hemoglobin
is proposed in section 2 B as well. Section 2 C describes the fit-
ting of the two free model parameters to the literature data for
the real part of the RI. In section 3, we apply the analysis to ex-
perimental data for oxygenated and deoxygenated hemoglobin,
respectively. The significance of the results is discussed and a
comparison to previous KK-analyses is made. In section 4, we
summarize our findings.
Terminology
Oxygenated hemoglobin is often abbreviated as "HbO2" while
deoxygenated hemoglobin is usually abbreviated as "Hb". In
order to simplify the notation we will use the term "Hb" for
hemoglobin in general, specifying the hemoglobin variant only
if relevant.
The RI of Hb solutions complex-valued. For the sake of
brevity, we refer to the real part of the RI by "real RI" and to
the imaginary part of the RI by "imaginary RI". This shall not
imply that the RI is purely real or imaginary.
2. MATERIALS AND METHODS
A. Absorption Spectra: Imaginary Part of Refractive Index
The Hb absorption spectra in the wavelength range
[250, 1100] nm reported in [8] were used in our calcula-
tions. These data were measured for Hb solutions produced
from human erythrocytes by repeated freezing and thawing fol-
lowed by centrifugation of the membrane components. These
homogeneous solutions did thus contain all the constituents of
human erythrocytes, except for the membranes (thickness less
than a few 10 nm, volume fraction less than a few % [6, 12, 13]).
We further use experimental data for pure Hb solutions to
supplement the absorbance data in the wavelength range
228 nm–250 nm [14].
Other researchers [9, 10] have previously relied on the data
compiled from various sources by Prahl [15]. However, since
not all of the compiled sources used human hemoglobin, these
data are not used here.
A.1. Literature Data
The literature absorption spectra we use in our calculations are
expressed in terms of the inverse absorption length µa(λ) at a
given mass concentration c or the molar extinction coefficient
10−3
10−5
)
1
−
g
L
(
/
)
λ
(
α
10−7
0
200
Applied Optics
2
100
oxyhemoglobin [8]
oxyhemoglobin [14]
deoxyhemoglobin [8]
deoxyhemoglobin [14]
generic peptide absorbance
water [16]
10−4
)
λ
(
O
2
H
κ
10−8
1000
400
800
vacuum wavelength λ/nm
600
Fig. 1. Imaginary refractive increment α(λ) of hemoglobin
in aqueous solutions (left axis, solid and dash-dotted lines).
Experimental data are taken from [8] (corrected for H2O ab-
sorbance) and [14]. A Lorentzian peak models the deep UV
absorbance of the peptide backbone (left axis, solid black line).
Imaginary RI κH2O(λ) of water (right axis, dashed line) [16].
Note that the quantities α, κH2O have different units and scal-
ing of y-axes. To compare the numerical values of α and κH2O,
one needs to multiply α by the respective concentration of the
RBC, e. g., cHb = 340 g L−1, cf. Eq. (5). Spline interpolation was
applied to obtain a step width of 1 nm.
ε M(λ), where λ is the vacuum wavelength of the light. The lat-
ter implies that the Lambert-Beer law holds, which is the case
for Hb at physiological (c ≈ 300 g L−1) or lower concentrations
[8]. Instead of µa(λ) or ε M(λ), we express the absorption spec-
tra in terms of the imaginary part of the complex RI
n(λ) = n(λ) + i κ(λ)
(1)
of the solution. Here n, κ are positive real functions. The con-
version rule is
κ(λ) =
µa(λ) λ
4π
=
ln 10 ε M(λ) c λ
4π M
.
(2)
The data on the molar extinction coefficient ε M(λ) from [14]
are converted to κ(λ), using the molar mass of the Hb tetramer
M = 64 458 g mol−1 [17].
A.2. Total Absorbance of Hb Solutions and Erythrocytes
In the spectral range of λ ∈ [250, 1100] nm the strongest absorp-
tion of light by RBCs and hence of blood is caused by Hb. Water
has a fairly low absorption coefficient in this region (cf. Fig. 1)
and other RBC components contained in the cytosol (e. g., other
proteins, sugars, ions) exhibit rather low concentrations. We
thus consider a two-component system, i. e.,
1. the solvent, or simply "Water" (H2O), occupying a volume
VH2O and
2. the absorbing solute, or simply "Hemoglobin" (Hb), de-
noting everything else contained in the erythrocyte cytosol
and occupying a volume VHb. The solute is treated as pure
Hb for its other physical properties, such as molar mass
and density.
Research Article
Applied Optics
3
Let the volume fraction occupied by Hb molecules be
φ = VHb/(VH2O + VHb) and the volume fraction occupied by
water molecules 1 − φ = VH2O/(VH2O + VHb).
Since the
Lambert-Beer law holds, we make the following ansatz for the
total imaginary RI (absorbance) of the solution
κ(λ) = φ κHb(λ) + (1 − φ) κH2O(λ),
(3)
where κH2O(λ) is the absorbance of pure water and κHb(λ) is
the absorbance of "pure hemoglobin in aqueous solution". It
should be noted that κHb is not equal to the imaginary RI of a
crystal of pure hemoglobin as is revealed by comparison with
data for the complex dielectric function of thin Hb films [18].
Eq. (3) is similar to the ansatz in [10], where, however, the
authors did not include the prefactor 1 − φ for the water ab-
sorbance, which results in a relevant difference as shown later.
This excluded volume effect, i. e., the fact that there is less water
in a solution of higher Hb concentration is essential and must
be taken into account since the volume fractions of Hb can be
as high as 26% or more.
The Hb volume fraction is related to the concentration by
B. Real Part of Refractive Index by Kramers-Kronig Relations
B.1. Literature Data
complete
experimental measurement of
the
The most
wavelength-dependent real RI of Hb solutions to date was pre-
sented in [8, 11]. Here, n(λ) was determined via measurements
of the spectral reflectance R(λ) at an interface between air and
Hb solution at normal incidence. The reflectance at normal
incidence is connected to the complex refractive index by the
Fresnel equation
R(λ) =(cid:12)(cid:12)(cid:12)(cid:12)
n(λ) − 1
n(λ) + 1(cid:12)(cid:12)(cid:12)(cid:12)
2
=
(n(λ) − 1)2 + κ(λ)2
(n(λ) + 1)2 + κ(λ)2 ,
(8)
which is easily solved for n(λ) when κ(λ) is known or when it
can be neglected because κ(λ) ≪ n(λ) − 1.
In [11] measurements at different concentrations were ana-
lyzed. A linear-affine dependence of the real RI on the concen-
tration was found and the result was expressed as
n(λ) = nH2O(λ) [1 + cHb β(λ)].
(9)
Here β(λ) is the concentration-specific increment of the real RI
relative to the water RI.
φ =
cHb
ρHb
,
(4)
B.2. Kramers-Kronig Relations
where ρHb is the mass density of a hypothetical solution con-
taining 100% Hb and 0% H2O. Both, κHb(λ) and ρHb are coeffi-
cients in a linear interpolation of spectra and density that holds
at least up to physiologically high concentrations. A value of
ρHb = 1330 g L−1 for proteins is given in [1], independent of
the type of protein. A normal physiological intra-erythrocyte
Hb concentration of 340 g L−1 thus corresponds to a volume
fraction of φ ≈ 0.26 or 26%.
Then the term for the absorbance contribution of Hb,
φ κHb(λ), becomes
φ κHb(λ) = cHb
κHb(λ)
ρHb
def. α(λ)
=
cHb α(λ),
(5)
where α(λ) is Hb's concentration-specific increment of the
imaginary RI, or the imaginary refractive increment. Since the Hb
concentration cHb is measured in g L−1, the unit for α is L g−1.
Eq. (3) becomes
κ(λ) = cHb α(λ) +(cid:18)1 −
cHb
ρHb(cid:19) κH2O(λ).
(6)
Eq. (6) together with Eq. (2) and Eq. (3) allows to compute
the imaginary refractive increment α(λ) from a measurement
of the inverse absorption length µ∗
a (λ) of a solution at known
concentration c∗
Hb as
α(λ) =
=
1
c∗
Hb
1
c∗
Hb
[κ∗(λ) + (φ∗ − 1) κH2O(λ)]
λ
4π(cid:20)µ∗
a (λ) +(cid:18) c∗
Hb
ρHb
− 1(cid:19) µa,H2O(λ)(cid:21) ,
(7)
where the asterisk ∗ denotes experimental data. The inverse ab-
sorption length of water µa,H2O(λ) is known to high accuracy
over a large spectral range λ ∈ [10 nm, 10 m] [16]. Hence this
formula allows to correct for the water absorption, which is im-
portant in the infrared (IR), where Hb absorbs only weakly.
The complex refractive index is a linear, causal response-
function to an incident wave. Hence its real and imaginary part
are related by KK relations. Expressed in terms of wavelengths,
the KK relations for the complex RI read
n(λ) − 1 = K[κ](λ) := −
κ(λ) = K−1[n](λ) = +
2
π
2
π
0
−Z ∞
−Z ∞
0
λ
λ
Λ
Λ2 − λ2 κ(Λ) dΛ,
λ
Λ2 − λ2 n(Λ) dΛ,
(10)
(11)
where Eq. (10) also defines the integral transform K in general.
The symbol := denotes a definition and the symbol −R denotes
the Cauchy principal value integral.
Applying Eq. (10) formally to the ansatz for the absorption
of the Hb solution Eq. (6), we obtain
n(λ) − 1 = cHb G(λ) +(cid:18)1 −
cHb
ρHb(cid:19) (nH2O(λ) − 1) ,
(12)
where G(λ)
:= K[α](λ) is the transformed spectrum (cf.
Eq. (10)). This formal transformation of the absorption results
in an equation for the real RI of the Hb solution
n(λ) = nH2O(λ) + cHb (cid:20)G(λ) −
nH2O(λ) − 1
ρHb
(cid:21)
(13)
def. B(λ)
= nH2O(λ) + cHb B(λ).
The linear-affine dependence of n(λ) on cHb in Eq. (13) is
in agreement with experimental findings [11] (cf. Eq. (9)). In
Eq. (6), we have formally split off the water absorption, such
that nH2O(λ) contributes to the background in the dispersion
relations of the Hb solutions (cf. Eq. (13)). Since the real RI
of water – unlike the real RI of Hb – is known to high accu-
racy, this provides valuable additional information compared
to the application of the KK-transform only to the measured
absorption spectrum of a Hb solution in the visible and near
UV/IR range that was presented in [9]. This idea was already
presented in [10]. However, due to our different ansatz for the
absorption, where we take into account the excluded water vol-
ume, we obtain a different result for B(λ) with an additional
Research Article
Applied Optics
4
term (nH2O(λ) − 1)/ρHb for the concentration dependence. We
discuss the differences between the results obtained in [9], [10]
and the result with our improved model in section 3 B.
For numerical values of nH2O(λ), we use a four-term Sell-
meier formula that is accurate to at least five decimal places
[19].
B.3. Additional Spectral Information Outside the Measured Range
The KK relations provide a formal tool to derive results like
Eq. (13). However, their application has a well known prob-
lem for the numerical evaluation: They are global integral trans-
forms that require the knowledge of real or imaginary RI at all
wavelengths λ ∈ [0, ∞[, which is practically impossible. Al-
though the integral kernel in Eq. (10) is decaying with increas-
ing distance from the pole at Λ = λ, it is long-ranged. Hence,
one cannot simply cut off the integration domain, i. e., use a fi-
nite dataset.
Water is transparent to visible light and its main regions of
absorption lie in the UV and the IR (Fig. 1). Due to these absorp-
tion bands water has a RI significantly different from 1, even
in the transparent regions and exhibits normal dispersion, i. e.,
n(λ) decreases with λ. This implies that neglecting such contri-
butions in the KK relations may lead to inaccurate results. We
will now describe the absorption features of Hb below 250 nm
by a mathematical model and include them in our expression
for the real RI.
In addition to the known absorption spectrum (Fig. 1) with
strong absorption in the vicinity of the Soret band at 420 nm,
Hb has an even stronger absorption peak in the deep UV. This
feature stems from the peptide bonds forming the backbone
of any polypeptide or protein, including the protein complex
hemoglobin and is characteristic for polypeptides and proteins.
The corresponding extinction coefficient curves ε(λ) are similar
among a variety of proteins and the absorbance maximum is
typically located at λ = 187 nm [20, 21].
This peptide-peak must be accounted for to perform a
proper KK analysis, but is, unfortunately, not resolved in the ex-
isting experimental Hb spectra. However, data were reported
for human and bovine albumin [21] – a protein found in blood
serum. Albumin is similar to hemoglobin in its mass and opti-
cal properties at wavelengths away from the characteristic Hb
absorption band at 420 nm. The absorption maximum for hu-
man albumin is reported as ε(187 nm) = 86.0 L g−1 cm−1, corre-
sponding to a value of α(187 nm) = 2.95 × 10−4 L g−1, which is
more than four times as high as the peak around 420 nm (Fig. 1).
We model this generic protein absorption using an anti-
symmetrized Lorentzian curve
αL(λ) = aL
1
π
Γ
(λ − L)2 + Γ2 − aL
1
π
Γ
(λ + L)2 + Γ2 ,
(14)
where Γ = 11.6 nm is the half width at half maximum of the
curve and L = 187 nm is the position of its maximum. This
model curve fulfills αL(−λ) = −αL(λ). This is important, as
the symmetries κ(−λ) = −κ(λ) and n(−λ) = n(λ) are implied
when using the KK relations in the form Eq. (10), Eq. (11) or the
analogous expressions for the frequency, where they are written
as an semi-infinite integral, denoted by the symbol −R ∞
As mentioned before, the deep UV spectrum and hence the
exact shape of the peptide absorption line is not available in the
literature. For proteins, Woods and O'Bar report that "the in-
crease in absorbance at 187 nm is threefold over that at 205 nm
and fourfold over that at 210 nm"[21]. This description fits well
to the half width of the curve of Γ = 11.6 nm. Going to even
0 .
lower wavelengths λ ≪ 187 nm there will be more absorption
features, since the inner electron shells of the atoms will be ex-
cited. We thus expect a variety of overlapping absorption lines
at these short wavelengths. On the other hand, this spectral re-
gion is fairly far away from the region of interest and the KK
relations contain a damping factor of 1/(λ2 − Λ2). Hence, the
exact line-shapes are practically irrelevant. For our model, it
suffices to add to the spectrum a delta-peak of unknown ampli-
tude located at zero wavelength, which accounts for the influ-
ence of extreme UV absorption by a constant offset in the real
RI: αδ(λ) = limλδ→0+
π
2 aδ λδ δ(λ − λδ).
In this respect, our approach is not different from [9, 10]: We
cannot predict the absolute value of the RI of Hb, but we need to
determine a constant by comparing to experimental data. How-
ever, we apply non-local fitting to optimize this constant, in-
stead of using just one single data point. Thus, our method is
more robust to uncertainties in both the absorption spectra and
the measured real RIs.
C. Fitting to Literature Values
With this model for the UV absorption, we have an absorption
spectrum
α(λ) = αlit(λ) + αL(λ) + αδ(λ),
(15)
where αlit(λ) represents the experimental literature data. For
the integral transform K (cf. Eq. (10)) the contribution from the
δ-peak αδ(λ) is Gδ(λ) = aδ and hence constant and the contri-
over the deep UV part of the spectrum, can be obtained ana-
bution GL(λ) = aL eGL(λ) from the Lorentzian, integrated only
lytically (see appendix B). Here eGL(λ) is the contribution for a
Lorentzian of unit amplitude. Thus, we are left with
G(λ) = Glit(λ) + aL eGL(λ) + aδ,
where only the first term Glit(λ) := K[αlit](λ) needs to be eval-
uated numerically.
Numerical evaluation is straightforward. We use an integra-
tion scheme, which evaluates the KK relations as a Riemann
sum with Taylor expansion at the singularities of the integrand
as described, e. g., in [22] and in appendix A. The experimen-
tal literature data αlit(λ) comprises two datasets: (1) the ab-
sorbance data in the range 228 nm–250 nm [14], which we refer
to as "ultraviolet" (UV) and (2) the data in the range 250 nm–
1100 nm [8], referred to here as "visible" (VIS). Although our
use of the terms "visible" and "ultraviolet" deviates from the
usual definition, it is convenient to distinguish between the two
datasets:
(16)
αlit(λ) =
αUV(λ) = data from [14] λ ∈ [228, 250] nm,
λ ∈ [250, 1100] nm,
αVIS(λ) = data from [8]
else.
0
(17)
The numerical KK transform is applied to both parts separately,
such that Glit(λ) = GUV(λ) + GVIS(λ).
C.1. Fitting of Free Parameters
Neither of the two free parameters of the model, aL and aδ can
be computed from literature data a priori with satisfying accu-
racy. For the peptide absorption aL, the order of magnitude
can be estimated from the semi-quantitative data [21], where
the absorbance maximum is given. It is important to keep in
mind that the KK transform of the peptide-peak in the deep UV
depends much stronger on the center position and the area un-
der the peak than on its actual maximum. Since the peak shape
Research Article
Applied Optics
5
is not quantitatively known, the peak height does not contain
enough information to determine aL.
The real RI
n(λ; aL, aδ) = nH2O(λ) + cHb B(λ; aL, aδ)
(18)
and the real refractive increment
B(λ; aL, aδ) = Glit(λ) −
nH2O(λ) − 1
+aL eGL(λ) + aδ
(19)
ρHb
{z
=:GH2O(λ)
}
linearly depend on the parameters aL and aδ. Thus we use a
linear least squares approach to optimize the parameter values.
The empirical model function in [11] is formally identical
to ours, but the quantity β(λ) = B(λ)/nH2O(λ) was consid-
ered instead of B(λ), cf. Eq. (9) and Eq. (13). The measure-
ments are given at wavelengths λi, i = 1, . . . , N and we de-
note these experimental values by β∗(λi). We convert them to
B∗
i = β∗(λi) nH2O(λi). In the following, the B∗
i will be referred
to simply as "the measurement data".
At each wavelength, the increment Bi := B(λi) consists of a
fixed part resulting from numerical KK transformation
B0,i = Glit(λi) + GH2O(λi)
(20)
and a function yet to be determined, which models the deep UV
contributions
fi = aLeG(λi) + aδ
def. hr(λ)
= ∑
r=L,δ
ar hr(λi).
(21)
Switching to matrix-vector notation, we write this function vec-
tor as f = H a, with H := {hr(λi)}ir ∈ RN×2 and a = (aL, aδ)T
is the parameter vector.
Now we want to minimize the deviation between KK results
with deep UV absorbance model B and measurement data B∗
B∗ − B = y − f ,
(22)
where y = B∗ − B0 is the data vector. The entries of vector B0
are given in Eq. (20). The linear least squares problem is then
χ2(a) → min with
χ2(a) := (y − f )T W (y − f )
(23)
=
1100 nm
∑
λi=250 nm
wij[B∗(λi) − B(λi; a)] [B∗(λj) − B(λj; a)],
where W = {wij}N
i,j=1 is a weight matrix given by the inverse
of the covariance matrix V of the data vector. The conditions
for minimal χ2(a) are solved by standard linear algebra, which
yields
C.2. Deoxyhemoglobin
Both, the absorption spectra of oxygenated and deoxygenated
hemoglobin are well known, cf. Fig. 1. However, the method
of [8] to measure the real RI could only be applied to oxyhe-
moglobin, whereas measurements with deoxyhemoglobin were
not possible due to precipitation resulting in backscattering of
light from highly concentrated solutions of deoxyhemoglobin.
Nevertheless, the KK analysis allows to derive a result for the
real refractive increment of deoxyhemoglobin. To this end, we
use the same model for the deep UV absorbance as above,
Lorentzian and delta-peak, with the coefficients aL, aδ found
by the least-squares fit for oxyhemoglobin. Combining this
with the KK transform of the spectrum αdeoxy(λ) for deoxyhe-
moglobin, we calculate the real refractive increment according
to
deoxy
Bdeoxy(λi) = B
0
deoxy
lit
= G
(λi) + f (λi)
(λi) + GH2O(λi) + f (λi).
(26)
The reason to simply use the same deep UV model curve for
oxyhemoglobin as well as for deoxyhemoglobin is that oxy-
genation affects the heme-groups in the hemoglobin complex.
The absorption peaks at about 420 nm and 560 nm are altered
by oxygenation (cf. Fig. 1) but not the peptide backbone caus-
ing the deep UV absorption.
D. Uncertainties
Both, the Kramers-Kronig transformation and the linear least-
squares fit to the data are linear transformations, which can for-
mally be carried out by matrix multiplication. Hence, it is easy
to perform the uncertainty propagation in terms of mean val-
ues and covariance matrices. A detailed description is given in
appendix C. All values for uncertainties given here correspond
to one standard deviation.
An important uncertainty contribution stems from the uncer-
tainty of the hemoglobin concentration cHb. Because this influ-
ences the concentration-specific transmittance and reflectance
spectra by a global factor, the uncertainties of the resulting real
refractive increment are strongly correlated between all wave-
lengths, i. e., the correlation coefficient
rij :=
cov(Bi, Bj)
qvar(Bi) var(Bj)
(27)
is close to +1 even if λi − λj is large. Here "cov" and "var"
denote covariance and variance, respectively.
In such a case
the use of the diagonal elements of the covariance matrix only,
i. e., the variances, may lead to unnecessarily less significant
results with respect to the uncertainty estimation. When, for
instance, the difference between two random variables X, Y is
considered, the variance is
var(X − Y) = var(X) + var(Y) − 2 cov(X, Y),
(28)
which may be small even if var(X), var(Y) are large.
a = arg min χ2(a) = (HT V−1H)−1HT V−1y
f = H a = H (HT V−1H)−1HT V−1
(24)
(25)
y
3. RESULTS AND DISCUSSION
=:F
{z
}
for parameter and function vector and B = B0 + f for the re-
fractive increment. The resulting amplitude of the Lorentzian
is shown in Fig. 1 as generic peptide absorbance.
A. Real Part of Refractive Index for Oxy- and Deoxyhe-
moglobin
Fig. 2 shows the result of the presented KK analysis for the
real refractive increment B(λ) for oxyhemoglobin along with
the experimental data B∗(λ) from [11]. The uncertainties are
Applied Optics
6
×10−4
oxyhemoglobin B(λ)
deoxyhemoglobin Bdeoxy(λ)
Research Article
×10−4
)
1
−
g
L
(
/
)
λ
(
B
3
2.8
2.6
2.4
2.2
KK result with UV model B
reflectance measurement B ∗ [11]
)
1
−
g
L
(
/
)
λ
(
B
3
2.8
2.6
2.4
2.2
300
400
500
600
700
800
900 1000 1100
300
400
500
600
700
800
900 1000 1100
vacuum wavelength λ/nm
vacuum wavelength λ/nm
Fig. 2. Real refractive increment B(λ) of aqueous solutions
of oxyhemoglobin: Experimental data [11] and results from
this work fitted to the data. The shaded bands indicate the
measurement uncertainties given in [11] and computed in ap-
pendix C, respectively.
shown as shaded bands, where the half-width of the band corre-
sponds to one standard deviation derived by covariance-matrix
calculus. The analogue result for deoxygenated hemoglobin is
shown in Fig. 3. Fig. 4 shows the derivative B′(λ) = d B(λ)/dλ.
Here the strong correlation is eliminated, resulting in much
It is evident from Fig. 4 that the main
smaller uncertainties.
uncertainty of the result concerns the exact vertical position of
the curve B(λ), whereas the uncertainties related to its shape,
described equivalently by the derivative B′(λ), are smaller.
The result B(λ) of the KK analysis for oxyhemoglobin and
the experimental data B∗(λ) as shown in Fig. 2 have a simi-
lar overall shape, but there are deviations which cannot be ac-
counted for by the measurement uncertainties of B∗(λ) given in
[11] and the uncertainties of the KK result derived as described
in appendix C. For instance, the difference between the peak at
433 nm and the dip at 399 nm for B∗ is
B∗(433 nm) − B∗(399 nm) = (7.5 ± 0.6) × 10−5 L g−1
while that for our result is
B(433 nm) − B(399 nm) = (5.47 ± 0.06) × 10−5 L g−1.
The small value for the uncertainty of B(433 nm) − B(399 nm)
is due to the fact that the strongly correlated concentration con-
tributions to the uncertainty cancel out almost completely for
this difference. This observation also holds for the derivative
B′(λ) (cf. Fig. 4). The difference between the two values clearly
exceeds the uncertainties.
The deviations between B(λ) and B∗(λ) can not be at-
tributed to unknown spectral absorptions outside the 250 nm–
1100 nm range, since features producing such discrepancies
would necessarily be inside this wavelength range. We have
modeled important spectral absorptions at the UV end of the
spectrum. Concerning possible IR absorptions that are not con-
sidered in the proposed model, we give the following notes:
The absorption spectra of aqueous hemoglobin solutions in the
IR between 1.1 µm and 2.6 µm are dominated by water [8] in-
dicating that the imaginary refractive increment of Hb is very
low in this region. Hypothetical absorption lines due to Hb at
Fig. 3. Real refractive increment of aqueous hemoglobin solu-
tions: Result for deoxyhemoglobin obtained by KK relations
with the same model for deep UV absorbance as for oxyhe-
moglobin. Cf. Fig. 2. Estimated standard deviations (shaded
bands) are larger for deoxyhemoglobin because Bdeoxy(λ) is
composed of more terms with uncorrelated concentration un-
certainties than B(λ).
even longer wavelengths would contribute to the real refrac-
tive index increment B(λ) in the form of constants or func-
tions with a gentle negative slope. Any possible influence of
the long-wavelength end of the spectrum of Hb would thus
change the agreement between our result B(λ) and the litera-
ture data B∗(λ) for the worse. Thus we conclude that no impor-
tant contribution to the absorption spectrum was missed at the
long-wavelength end. The KK relations themselves are valid as
long as the framework within which the data were measured
holds, i. e., classical electrodynamics and linear, causal media.
Optical activity of Hb has been examined previously [14, 23]
and can be ruled out as well, since the expected effect is too
weak. Furthermore, we can exclude biological variability, since
absorption and reflectance/refraction data were measured on
the same type of sample [8, 11]. This implies that the absorption
and refraction data presented in [8, 11] are not self-consistent
with the given measurement uncertainties.
Lastly, we address the question, whether these discrepancies
are of practical importance. At λ = 399 nm, where the discrep-
ancy is largest, one has B∗ = 2.24 × 10−4 L g−1 from [11] and
B = 2.41 × 10−4 L g−1 for our analysis. At a concentration of
cHb = 340 g L−1 the resulting real RIs are n∗ = nH2O + cHb B∗ =
1.420 and n = nH2O + cHb B = 1.426, respectively and the resid-
ual is n − n∗ = 0.006 = 0.4% n∗. Since this value is in the sub-
percent region it may, at first glance, seem unimportant whether
one uses n∗ or n. However, when RBCs are examined with op-
tical methods such as phase contrast microscopy or light scat-
tering in a flow cytometer, they are usually immersed in wa-
ter with an RI of nH2O(399 nm) = 1.344, or a saline solution of
slightly higher RI. Hence the complex contrast between cell and
surrounding medium is
m − 1 :=
n
nH2O
− 1 =(0.061 + 0.008i
0.057 + 0.008i our result
[11]
.
(29)
In phase contrast microscopy, the signal for the optical thick-
ness of the cell is directly proportional to the contrast m − 1. It
Research Article
Applied Optics
7
oxyhemoglobin d
dλ
deoxyhemoglobin d
B(λ)
dλ Bdeoxy(λ)
×10−6
)
1
−
m
n
1
−
g
L
(
/
)
λ
(
B
′
6
4
2
0
-2
300
400
500
600
700
800
900 1000 1100
vacuum wavelength λ/nm
Fig. 4. Derivative of the real refractive increment B′(λ) =
d
dλ B(λ): Comparison between results for deoxyhemoglobin
and oxyhemoglobin. Estimated standard deviations (shaded
bands, barely visible) of B′(λ) are much smaller than for B(λ)
(Fig. 3). This is because most of the (strongly correlated) con-
centration uncertainty cancels out in the derivative.
is also a parameter governing the light scattering by cells. The
Rayleigh-Gans-Debye theory is an approximate description of
light scattering for the limiting case of particles with low con-
trast and moderate size. It predicts a scattered electric field pro-
portional to m − 1. The real parts of the above two values for
m − 1 differ by more than 7%, indicating that any quantitative
analysis of the interaction of RBCs with light that requires a pri-
ori knowledge of the contrast can not be more accurate than
this.
B. Comparison to Previous KK Analyses
We will now briefly review the previous investigations employ-
ing KK relations [9, 10] and compare the methods. In [9], the
authors started from Eq. (10) and applied it to the Hb absorp-
tion spectra in a finite spectral range, i. e., instead of Eq. (6) they
assumed
κ[9](λ) = cHb αVIS(λ) = cHb( α(λ) λ ∈ [a, b]
else
0
,
(30)
where [a, b] is the measured spectral range, i. e., here [a, b] =
[250, 1100] nm. The authors then used a subtractive form of the
KK relations, where the difference n(λ) − n(λ0) is considered
which yields
n[9](λ) = n(λ0) + cHb [GVIS(λ) − GVIS(λ0)].
(31)
Here, GVIS(λ) := K[αVIS](λ) is the dispersion resulting from
the measured spectrum. The free parameter n(λ0) is fixed by a
refractometric measurement at wavelength λ0 = 800 nm. If the
non-subtractive KK relations had been used instead, the result
would have been
n(λ) = 1 + cHb GVIS(λ),
(32)
which is off the true value by a significant amount. One can
remove this discrepancy by replacing the 1 in the above expres-
sion by a free parameter, which can be interpreted as deep UV
absorption. Again, this free parameter can be fixed by a single
experiment n∗(λ) [11]
this paper
[9]
[10]
)
λ
(
n
x
e
d
n
i
e
v
i
t
c
a
r
f
e
r
1.46
1.44
1.42
1.4
400
600
800
1000
vacuum wavelength λ/nm
Fig. 5. Comparison between the experimental real RI derived
from measured refractive increment of Hb solutions [11] and
the water RI [19] (measurement uncertainty indicated by
shaded band) and different Kramers-Kronig analyses. Con-
centration is cHb = 287 g L−1. The method in [9] ignores ab-
sorption outside the measured range and in [10] the UV and
IR absorption of water was taken into account. In the present
work the excluded water fraction was accounted for, as well
as a model for deep UV absorbance. All KK methods were ap-
plied to the same dataset depicted in Fig. 1. Curves calculated
according to [9, 10] are matched to the experimental curve at
λ0 = 800 nm. The two free parameters characterizing the deep
UV model in our method are determined by a global fit to the
experimental refractive increment.
measurement at λ0. The result is then the same as in Eq. (31).
However, the subtractive KK transform K[α](λ) − K[α](λ0) can
be re-written into a single integral where the kernel decays
faster than in the standard KK relations, which is numerically
favorable and thus given as an argument for the use of subtrac-
tive relations. Thus, the subtractive form of KK relations masks
the lack of knowledge of the absorption spectrum outside the
measured spectral range: At least one important absorption
peak at short wavelengths has apparently been omitted. As
long as the missing peak is far away from the region of inter-
est, the simple addition of a constant works fine. However, if
the location of the peak becomes important as is the case for
water, this model is insufficient.
In [10], the authors made the ansatz
κ[10](λ) = κH2O(λ) + cHb αVIS(λ).
(33)
for the imaginary RI of the Hb solution, which takes the absorp-
tion due to water into account but neglects the excluded volume
effect for water in contrast to our approach in Eq. (6). Apart
from this difference, the formal application of the KK relations
in the present work is identical to that in [10]. The result was
n[10](λ) = nH2O(λ) + cHb GVIS(λ),
(34)
which provides also a theoretical derivation of the empirical
finding n(λ) = nH2O(λ)[1 + cHb β(λ)] reported in [11]. How-
ever, the result that β(λ) = GVIS(λ)/nH2O(λ) is incomplete, as
we have discussed. Subtractive KK relations were used as well
to match the RI at λ0 = 800 nm.
To compare these two previously presented methods with
ours, we applied them to the spectra of oxyhemoglobin pre-
Research Article
Applied Optics
8
×10−3
)
λ
(
∗
n
−
)
λ
(
n
l
a
u
d
i
s
e
r
x
e
d
n
i
e
v
i
t
c
a
r
f
e
r
5
0
-5
-10
-15
-20
this paper
[10]
400
600
800
1000
vacuum wavelength λ/nm
Fig. 6. Residuals between calculated real RI n(λ) and experi-
mental real RI n∗(λ) [11] for our method and the method ap-
plied in [10], cf. Fig. 5.
sented in [8] and shown in Fig. 1. As an example, a con-
centration of cHb = 287 g L−1 was assumed and n(λ0 =
800 nm) = 1.403 was taken from [8] as well. The compari-
son of the two methods applied in [9, 10] with our method
and the experimental dispersion curve given in [11] is shown
in Fig. 5. Neglecting the water influence as in [9] yields a dis-
persion curve which substantially deviates from the measure-
ment (dash-dotted black line in Fig. 5) everywhere, except at
and above λ0, where it was matched. With the water absorp-
tion and the resulting dispersion taken into account [10] the
agreement with the experimental result is already much better.
However, the result from [10] and experimental curve substan-
tially differ in the UV below 375 nm. The agreement in the UV
is much better in our approach. This is also evident from Fig. 6,
where the residuals between KK computations and experiment
are plotted.
4. SUMMARY
The complex refractive index of a hemoglobin solution, which
forms the cytoplasm of erythrocytes and determines their opti-
cal properties can be computed as
n(λ) = nH2O(λ) + i κH2O(λ) + cHb[B(λ) + i α(λ)],
(35)
where κH2O(λ) is negligibly small for λ ∈ [250, 1100] nm and
physiological hemoglobin concentrations cHb.
In the present
work, we have computed the real refractive increment B(λ)
from experimental spectra of the imaginary refractive incre-
ment α(λ) for λ ∈ [250, 1100] nm. We formally separated the
solution's imaginary RI into a water and a hemoglobin part
and then applied the Kramers-Kronig relations to obtain the
real RI and thus an expression for B(λ), Eq. (13). The absorp-
tion spectra available in the literature [8, 14] do not resolve
the strong UV absorbance of hemoglobin's peptide-backbone.
Hence we modeled it by a Lorentzian line of unknown am-
plitude, which introduces a free parameter aL into the expres-
sion for B(λ). A second free parameter aδ is introduced as a
wavelength-independent term accounting for extreme UV ab-
sorbance, cf. Eq. (19). These two free parameters were deter-
mined by a linear least squares fit to the literature data B∗(λ)
[11], the result of the fit is denoted by B(λ).
We evaluated spectra for oxygenated and deoxygenated
hemoglobin. Data files of the results are provided as supple-
mentary material for B(λ) (Data File 1), Bdeoxy(λ) (Data File 3),
the corresponding covariance matrices (Data Files 2 and 4), the
converted literature data for α(λ) and αdeoxy(λ) [8] and the val-
ues for nH2O(λ) computed according to [19] (Data Files 1 and 3)
as well as for B′(λ) (Data File 5).
The uncertainties for the curve B(λ) were computed and
reveal that its shape is resolved much more accurately than
in the measurements B∗(λ) in [11]. The analysis furthermore
shows that the real refractive increments for oxygenated and de-
oxygenated hemoglobin differ significantly from each other for
wavelengths between 350nm and 600nm. In the vicinity of the
absorption band at 420nm, deviations between our result for
the real refractive increment and the experimental values [11]
clearly exceed the measurement uncertainties, where the curve
obtained by our KK analysis has a much smoother shape and
does not exhibit the non-monotonic up-and-down movements
for wavelengths larger than 600 nm of the data presented in
[11].
A. APPENDIX: NUMERICAL INTEGRATION SCHEME
FOR KRAMERS-KRONIG RELATIONS
For numerical integration of KK relations, we follow the con-
cept described in [22]: a Riemann sum with Taylor expansion
at the singularities of the integrand. Numerical stability was
tested by comparing to the analytical transformations of differ-
ent Lorentzian and rectangular profiles.
We want to evaluate numerically the expression
λ
Λ
π G(λ) = −2−Z λb
λa (cid:18) 2
= −Z λb
λa
Λ
λ
Λ2 − λ2 α(Λ) dΛ
−
1
Λ + λ
−
1
Λ − λ(cid:19) α(Λ) dΛ.
The measurement data are given at discrete points
αi := α(λi),
λi := λa + t(cid:18)i −
1
2(cid:19) ,
i = 1, . . . , N,
(36)
(37)
(38)
where λ1 = 250 nm, λN = 1100 nm and t = 1 nm. Hence
λa = 249.5 nm, λb = 1100.5 nm and N = 851. We only eval-
:= G(λi). The third
uate the integral at the grid points Gi
term in the integral has a singularity at λ = Λ. The first two
terms are not singular, hence no principal value integrals have
to be used here. All integrals for non-singular integrands are
approximated by Riemann sums, including the third term for
Λ /∈ [λi − t/2, λi + t/2]. The remaining principal value integral
can be re-written by Taylor series expansion of the integrand.
Using only the lowest non-vanishing order yields
Numerically, we use the nearest-neighbor lattice-derivatives
1
λj − λi
N
∑
j6=i
j=1
αj t − t α′
i.
(39)
1 < i < N
i = 1
i = N
.
(40)
π Gi ≈
1
−
λj
N
∑
j=1 2
λj + λi! αj t −
i =
(αi+1 − αi−1)/2
α2 − α1
αN − αN−1
t α′
Note that Eq. (39) can also be written as G = K α, where K is a
N × N matrix.
Research Article
Applied Optics
9
B. APPENDIX: UV
DOMAIN
KRAMERS-KRONIG-
TRANSFORM OF LORENTZIAN CURVE
If the KK transform (Eq. (10) is applied to the antisymmetric
Lorentzian line in the wavelength domain Eq. (14) the result is
2. Detector/instrument noise in the reflectance spectra [8,
11], resulting in a wavelength-independent uncertainty of
3 × 10−6 L g−1 [11] for the real refractive increment β∗(λ).
3. Uncertainties of the solutions' Hb concentration cHb.
(41)
4. Uncertainty of the hemoglobin density ρHb relating mass
concentration cHb and volume fraction φ. Here we assume
one digit, i. e., uρHb = 10 g L−1 = 0.75% ρHb.
K[αL](λ) = aL
1
π(cid:18)
π(cid:18)
1
− aL
λ − L
(λ − L)2 + Γ2 +
λ + L
(λ + L)2 + Γ2 −
L
L2 + Γ2(cid:19)
L2 + Γ2(cid:19) .
L
In our work, however, we need to integrate only over the deep
ultraviolet spectrum below a threshold λa. The analytical ex-
pression for this reads
GL(λ) =
2
Γ
1
1
π
−
Λ
=
aL
(λ − L)2 + Γ2 +
−λa(cid:18) 1
Λ − λ(cid:19) αL(Λ) dΛ
−Z λa
π2(cid:26) 1
(λa + L)2 + Γ2(cid:19)
ln(cid:18) (λa − L)2 + Γ2
×(cid:20)
λa + λ(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) ×(cid:20)
− ln(cid:18)(cid:12)(cid:12)(cid:12)(cid:12)
+(cid:18)π − arctan
×(cid:20)
(λ − L)2 + Γ2 −
− arctan
λa − λ
λa − L
λ − L
λ + L
Γ
Γ
(λ − L)2 + Γ2 −
Γ
Γ
λa + L(cid:19)
(λ + L)2 + Γ2 +
(λ + L)2 + Γ2 −
2Γ
L2 + Γ2(cid:21)
(λ + L)2 + Γ2(cid:21)
Γ
2L
L2 + Γ2(cid:21)(cid:27)
5. Uncertainty of the complex RI of water. This influence is
negligible.
These uncertainties can be divided into two classes, separat-
ing 1.–2. from 3.–4.:
(i) Noise in the measured reflectance and transmittance spec-
tra, affecting the spectral data locally. The corresponding
quantities are labeled with the subscript "noise". As a
model, we assume white noise, i. e., two measurement er-
rors at different wavelengths are not correlated.
(ii) Measurement uncertainties in scalar quantities (cHb, ρHb)
occurring as prefactors in the spectra. These affect the
spectra by a global factor, i. e., the errors at different wave-
lengths have correlation coefficient +1. The corresponding
quantities are labeled with a subscript "conc".
In addition there are model errors, which arise due to un-
known absorption spectra outside the measured range and
which we discuss in appendix E.
(42)
for λ > λa. Eq. (42) is useful to describe the deep UV part of
the absorption spectrum without the need for numerical inte-
gration.
We also define the transformation for unit amplitude eGL(λ),
such that GL(λ) = aL eGL(λ).
C. APPENDIX: DETAILS OF UNCERTAINTY PROPAGA-
TION
A. Uncertainties of Input Data
We now recapitulate how spectral data are measured, in order
understand which contributions to the combined uncertainties
occur in the problem at hand.
The absorption spectra (inverse absorption length µa(λ),
imaginary RI κ(λ)) are measured via the attenuation of a beam
of light passing through a sample of known thickness. This is
performed for a number of wavelengths λi, i = 1, . . . , N. The
molar extinction coefficient ε M(λ) and the imaginary refractive
increment α(λ) are obtained from these quantities by dividing
with the separately determined molar/mass concentration of
Hb in the solution.
In [8, 11] spectral reflectance measurements were performed
and evaluated using the Fresnel equation Eq. (8) to obtain the
real RI n(λ). To determine the experimental value of the real
refractive increment B∗(λ), several curves at different concen-
trations cHb were recorded and the slope of n(λ) with respect
to cHb was computed for all λ and normalized to the known
water RI to obtain β∗(λ).
Having this in mind, measurement uncertainties of the fol-
lowing quantities need to be taken into account:
1. Detector/instrument noise in the absorption spectra. This
affects the inverse absorption length µa(λ) in [8] and the
molar extinction coefficient ε M(λ) in [14].
B. Error Model and Uncertainty Propagation
For any random vector ξ ∈ RN and any non-random linear
transform A ∈ RN×N, we have for η = A ξ
(43)
(44)
(45)
E(η) = A E(ξ)
V(η) = A V(ξ) AT,
where E denotes the expectation value (or mean) and
[V(ξ)]ij = cov(ξi, ξ j) = E(cid:16)[ξi − E(ξi)] [ξ j − E(ξ j)](cid:17)
denotes the covariance matrix. This is independent of the un-
derlying probability distribution, only implying that the first
and second moments exist. We restrict our uncertainty analy-
sis to mean values and covariance matrix, which describes the
corresponding probability distributions fully only in the special
case of a normal distribution.
The uncertainties of the literature absorption spectra are not
given; neither the noise nor the concentration uncertainties are
quantified. We thus estimate and model them as follows. The
covariance matrix for the absorption spectra is assumed to be
V(α) = Vnoise(α) + Vconc(α). Here
Vnoise(α) = σ2
noise,rel diag(α)2
(46)
is the local/uncorrelated part due to (white) detector noise.
This matrix is diagonal.
Vconc(α) = σ2
conc,rel α αT
(47)
is the global/correlated part due to concentration uncer-
tainty. This matrix has a tensor-product structure. We use
σnoise,rel = 0.5% for the relative noise level and σconc,rel = 1% for
the relative concentration uncertainty [24].
Research Article
Applied Optics
10
Eq. (47) is motivated as follows: Concentration errors change
a spectrum by a global factor, such that the measured value
αmeas = (1 + ξ)α,
(48)
is off the (unknown) true value α. We assume this error to be
unbiased, i. e., E(ξ) = 0 andpvar(ξ) = σconc,rel corresponds
to the relative concentration uncertainty of the solution.
E(αmeas) = α, and
I. e.,
V(αmeas) = var(ξ) α αT = var(ξ)
=σ2
conc,rel αmeas αT
≈ σ2
E(αmeas) E(αmeas)T
(49)
{z }
meas,
conc,rel
which is Eq. (47). In the following we will drop the distinction
between unknown true values and measured values and substi-
tute the latter for the former.
This error model is formally invariant under linear transfor-
mation. The measured value of some quantity obtained by a
linear transform A (e. g., the KK transform)
ηmeas = A αmeas = (1 + ξ)A α = (1 + ξ) η,
(50)
is formally identical to Eq. (48) such that the variance V(ηmeas)
can be computed from E(ηmeas) in the same manner as for the
original quantity α, i. e., V(ηmeas) = var(ξ)η ηT.
C. Uncertainty Propagation in Kramers-Kronig Relations
For discrete data points, the KK relations can be written in ma-
trix vector form, i. e., G = K α, where K is a N × N matrix (see
appendix A). If V(α) is the (co-)variance matrix of the spectrum
α, then the covariance matrix of the transform G is
V(G) = K V(α) KT
and with the model for V(α) one obtains
V(G) = Vnoise(G) + Vconc(G),
(51)
(52)
conc,rel G GT has the same tensor prod-
where Vconc(G) = σ2
uct structure as Vconc(α) in Eq. (47). In contrast, Vnoise(G) =
K Vnoise(α) KT is, unlike Vnoise(α), not diagonal, because of the
non-locality of the KK transform.
D. Uncertainty Propagation in Linear Fit
The covariance matrix V(y) of
y = B∗ − B0 = B∗ − GVIS − GUV − GH2O
(53)
is obtained by
V(y) = V(B∗) + V(B0)
= V(B∗) + V(GVIS) + V(GUV) + V(GH2O).
(54)
This decomposes into a noise and a concentration term V(y) =
V(y)noise + V(y)conc. For the contributions from concentration
uncertainties, one finds (cf. Eq. (54) and Eq. (49))
Vconc(y) =
4
∑
j=1
vj vT
j .
with
v1 := σconc,rel B∗,
v2 := σconc,rel GVIS,
v3 := σconc,rel GUV,
v4 :=
uρHb
ρHb
GH2O.
(55)
(56)
Similarly, we obtain Vconc(B∗) = v1 vT
1 .
However, for weighting the linear fit, we do not use this co-
variance matrix, but only the noise terms, i. e., V = Vnoise(y)
with
Vnoise(y) = Vnoise(B∗) + K Vnoise(αVIS) KT + K Vnoise(αUV) KT
(57)
(noise in nH2O is negligible). Here Vnoise(B∗) is a diagonal ma-
trix containing the uncertainties given for the refractive incre-
ment in [11]. The reason to use V = Vnoise(y) is that the tensor-
product structure (cf. Eq. (55) and Eq. (49)) of the systematic
covariance matrices makes them singular and hence the full co-
variance matrix (including noise) close to singular, which is a
problem for the numerics. But after all, weighting with uncer-
tainties that arise from a global prefactor for all values has little
sense.
Since f = F y (Eq. (25)) is obtained by linear transformation,
the covariance matrix is formally obtained as
Vnoise( f ) = F V FT = H (HT V−1H)−1HT.
For the concentration contributions one obtains
Vconc( f ) =
4
∑
j=1
wj wT
j with wj = F vj.
(58)
(59)
Although f is formally the result of the fit, the quantity we
are interested in is not f but rather
B = B0 + f = B0 + F y = B0 + F (B∗ − B0) = (F − 1) B0 + F B∗,
(60)
where 1 denotes the identity matrix. We can assume B∗ (the
measured refractive increment) and B0 (the part of the refrac-
tive increment computed from absorption spectra and water RI)
to be uncorrelated and obtain we obtain
V( B) = (F − 1) V(B0) (FT − 1) + F V(B∗) FT
= V(B0) + V( f ) − F V(B0) − V(B0) FT.
(61)
This also decomposes into
V( B) = Vnoise( B) + Vconc( B),
(62)
where both parts are computed separately. Vconc( B) contributes
stronger to the diagonal elements of V( B) (i. e., the variances of
the Bi) than Vnoise( B). On the other hand,Vconc( B), is strongly
correlated among all elements, whereas Vnoise( B) is not. To
illustrate the degree of correlation, the numerical derivative
B′(λi), obtained from finite differences, is plotted in Fig. 4 with
the corresponding standard deviations.
D.1. Deoxyhemoglobin
The refractive increment for deoxyhemoglobin in Eq. (26) can
also be written as
Bdeoxy = G
− F Glit − (F − 1) GH2O + F B∗,
(63)
deoxy
lit
from which follows
deoxy
Vnoise(Bdeoxy) = Vnoise(G
lit
)
+ F Vnoise(Glit) FT + F Vnoise(B∗) FT
6
∑
j=1
uj uT
j
Vconc(Bdeoxy) =
with
uj := F vj for j = 1, 2, 3;
deoxy
VIS
u5 := σconc,rel G
,
u4 := (F − 1) v4
u6 := σconc,rel G
deoxy
UV .
(64)
(65)
(66)
Research Article
Applied Optics
11
13. V. Heinrich, K. Ritchie, N. Mohandas, and E. Evans, "Elastic thickness
compressibilty of the red cell membrane," Biophysical Journal 81, 1452–
1463 (2001).
14. Y. Sugita, M. Nagai, and Y. Yoneyama,
"Circular dichroism of
hemoglobin in relation to the structure surrounding the heme," Journal
of Biological Chemistry 246, 383–388 (1971).
absorption
hemoglobin,"
"Optical
15. S.
Prahl,
of
http://omlc.org/spectra/hemoglobin/ (1999).
16. D. J. Segelstein, "The complex refractive index of water," Ph.D. thesis,
University of Missouri – Kansas City (1981).
17. G. Braunitzer, "The molecular weight of human haemoglobin." Biblio-
theca haematologica 18, 59 (1964).
18. H. Arwin, "Optical properties of thin layers of bovine serum albumin, γ-
globulin, and hemoglobin," Applied Spectroscopy 40, 313–318 (1986).
19. M. Daimon and A. Masumura, "Measurement of the refractive index
of distilled water from the near-infrared region to the ultraviolet region,"
Applied Optics 46, 3811–3820 (2007).
20. A. R. Goldfarb, L. J. Saidel, and E. Mosovich, "The ultraviolet absorp-
tion spectra of proteins," Journal of Biological Chemistry 193, 397–404
(1951).
21. A. H. Woods and P. R. O'Bar, "Absorption of proteins and peptides in
the far ultraviolet," Science 167, 179–181 (1970).
22. C. A. Emeis, L. J. Oosterhoff, and G. d. Vries, "Numerical evaluation of
Kramers-Kronig relations," Proceedings of the Royal Society of London.
Series A, Mathematical and Physical Sciences 297, 54–65 (1967).
23. L. H. Laasberg and J. Hedley-Whyte, "Optical rotatory dispersion of
hemoglobin and polypeptides: Effect of halothane," Journal of Biologi-
cal Chemistry 246, 4886–4893 (1971).
24. K. Witt, H. U. Wolf, C. Heuck, M. Kammel, A. Kummrow, and
J. Neukammer, "Establishing traceability of photometric absorbance val-
ues for accurate measurements of the haemoglobin concentration in
blood," Metrologia 50, 539 (2013).
E. Peptide Absorption: Influence of Peak Shape
We have assumed the exact shape of the peptide peak to be
of minor importance, such that the results depend mainly on
its position, total strength and width.
In order to justify this
assumption quantitatively, we evaluated an alternative model,
where the peak has a rectangular shape.
The comparison between results with the Lorentzian peak
centered at L = 187 nm and half width at half maximum
Γ = 11.6 nm and a rectangle centered around LΠ = L and half
width ΓΠ = Γ reveals that deviations between the two mod-
els are negligible in comparison to the propagated data uncer-
tainties for the majority of wavelengths. Only at wavelengths
λ < 400 nm does the deviation B(λ) − BΠ(λ) exceed the esti-
mated uncertainties due to noise, but is still smaller than the to-
tal uncertainty including concentration errors. For λ > 400 nm
the deviation is at least one order of magnitude smaller than the
total uncertainties.
ACKNOWLEDGMENT
The authors would like to thank Moritz Friebel for providing in
tabulated form the absorbance data for oxygenated and deoxy-
genated hemoglobin published in [8] as figures.
REFERENCES
1. R. Barer and S. Joseph, "Refractometry of living cells: Part I. basic
principles," Quarterly Journal of Microscopical Science s3-95, 399–423
(1954).
2. Y. Park, T. Yamauchi, W. Choi, R. Dasari, and M. S. Feld, "Spectro-
scopic phase microscopy for quantifying hemoglobin concentrations in
intact red blood cells," Opt. Lett. 34, 3668–3670 (2009).
3. Y. Sung, N. Lue, B. Hamza, J. Martel, D. Irimia, R. R. Dasari, W. Choi,
Z. Yaqoob, and P. So, "Three-dimensional holographic refractive-index
measurement of continuously flowing cells in a microfluidic channel,"
Physical Review Applied 1, 014002 (2014).
4. E. Schonbrun, R. Malka, G. Di Caprio, D. Schaak, and J. M. Hig-
gins, "Quantitative absorption cytometry for measuring red blood cell
hemoglobin mass and volume," Cytometry Part A 85, 332–338 (2014).
5. Y. Kim, H. Shim, K. Kim, H. Park, S. Jang, and Y. Park, "Profiling in-
dividual human red blood cells using common-path diffraction optical
tomography," Scientific Reports 4, 6659 (2014).
6. D. H. Tycko, M. H. Metz, E. A. Epstein, and A. Grinbaum, "Flow-
cytometric light scattering measurement of red blood cell volume and
hemoglobin concentration," Applied Optics 24, 1355–1365 (1985).
7. N. Mohandas, Y. R. Kim, D. H. Tycko, J. Orlik, J. Wyatt, and W. Groner,
"Accurate And Independent Measurement Of Volume And Hemoglobin
Concentration Of Individual Red-Cells By Laser-Light Scattering," Blood
68, 506–513 (1986).
8. M. Friebel and M. Meinke,
the complex refrac-
tive index of highly concentrated hemoglobin solutions using transmit-
tance and reflectance measurements," Journal of Biomedical Optics 10,
064019–064019–5 (2005).
"Determination of
9. D. J. Faber, M. C. G. Aalders, E. G. Mik, B. A. Hooper, M. J. C. van
Gemert, and T. G. van Leeuwen, "Oxygen saturation-dependent ab-
sorption and scattering of blood," Physical Review Letters 93, 028102
(2004).
10. O. Sydoruk, O. Zhernovaya, V. Tuchin, and A. Douplik, "Refractive
index of solutions of human hemoglobin from the near-infrared to the ul-
traviolet range: Kramers-Kronig analysis," Journal of Biomedical Optics
17, 115002–115002 (2012).
11. M. Friebel and M. Meinke, "Model function to calculate the refractive
index of native hemoglobin in the wavelength range of 250–1100 nm
dependent on concentration," Applied Optics 45, 2838–2842 (2006).
12. H. J. Deuling and W. Helfrich, "Red blood cell shapes as explained on
the basis of curvature elasticity." Biophysical Journal 16, 861 (1976).
|
1303.4269 | 1 | 1303 | 2013-03-15T17:08:17 | Cohesion-decohesion asymmetry in geckos | [
"physics.bio-ph",
"cond-mat.soft"
] | Lizards and insects can strongly attach to walls and then detach applying negligible additional forces. We propose a simple mechanical model of this phenomenon which implies active muscle control. We show that the detachment force may depend not only on the properties of the adhesive units, but also on the elastic interaction among these units. By regulating the scale of such cooperative interaction, the organism can actively switch between two modes of adhesion: delocalized (pull off) and localized (peeling). | physics.bio-ph | physics |
Cohesion-decohesion asymmetry in geckos
1Dipartimento di Scienze dell'Ingegneria Civile e dell'Architettura (ICAR), Politecnico di Bari, Italy and
G. Puglisi1 and L. Truskinovsky2
2LMS, Ecole Polytechnique, 91128, Palaiseau, France
Lizards and insects can strongly attach to walls and then detach applying negligible additional forces.
We propose a simple mechanical model of this phenomenon which implies active muscle control. We
show that the detachment force may depend not only on the properties of the adhesive units, but also
on the elastic interaction among these units. By regulating the scale of such cooperative interaction,
the organism can actively switch between two modes of adhesion: delocalized (pull off) and localized
(peeling).
PACS numbers: 68.35.Np, 87.17.Rt, 87.85.gj, 87.10.Pq., 87.10.Hk, 62.20.Qp, 71.15.Nc
I.
INTRODUCTION
Active mechanisms involved in biological adhesion in
living systems are of broad theoretical interest in view of
potential applications in bio-inspired adhesion devices.
One of the most challenging issues concerns the reconcil-
iation of strong adhesion [1] with easy detachment [2, 3].
Experimental studies reveal that biological adhesion
at the organismic level is typically mediated by fibrillar
microstructures which ensure a molecular level contact
in the presence of surface roughness. The necessity of
avoiding clustering entails a hierarchy of adhesion de-
vices spanning a wide range of scales [4, 5]. In the case
of geckos the macroscopic adhesion force at the level of a
foot results from smaller forces at the scale of individual
pads, each composed of tens of lamellae which in turn
incorporate hundreds of thousands of setae. Each seta
is split into hundreds of spatula shafts ending with spat-
ula pads at the sub-micrometer scale where the adhesive
force is ultimately generated by van der Waals forces [3].
Important insights into the functioning of microfibril-
lar adhesive devices in geckos were obtained from AFM
(Atomic Force Microscopy) experiments at the scale of
spatulas [6] and setae [7, 8] and from attempts to ar-
tificially microfabricate fibrillar microstructures [9].
In
particular, these studies revealed a strong dependence
of the adhesive forces on the angle formed by the se-
tal shafts with the adhesion surface, which suggests that
easy detachment may be achieved by active reorientation
of the single microscopic fibers. Theoretical understand-
ing of the angle dependence of adhesion at the level of a
single spatula is mostly based on the study of Kendall's
model [2] which has been recently generalized to account
for the asymmetric attachment-detachment behavior of a
single seta [10] and for fiber tilting [11]. The main idea is
that decohesion can be described as peeling which implies
that a Griffith's fracture takes place in an infinitely local-
ized tip of a steadily propagating crack. An alternative
model suggested in [7, 8] links the cohesion-decohesion
asymmetry with a dependence of the cohesive strength
on the tangential component of the force. An interesting
attempt to reconcile such friction-based approach with
Kendall's fracture model was proposed in [12] where a fi-
nite prestretch in the adhering layer was used to control
the critical angle of adhesion.
While fiber reorientation is clearly important for gecko
adhesion, we propose in this paper a complementary
mechanism that can be broadly characterized as the pos-
sibility of active switching between localized (peeling)
and delocalized (pulling off) fracture. We argue that
the organism can control the modality of detachment by
changing the level of coupling among individual fibrillar
agents [13]. In contrast to Kendall's model, based on the
assumption that the adhering layer can support only in-
plane forces, we assume that this layer can have a shear
stiffness which mimics bending resistance and is respon-
sible for the cooperative effects. We also assume that the
organism can actively switch between the regimes of high
and low stiffness depending on the force that has to be
exerted. We discuss potential mechanisms of how such
control at different scales of the fibrillar micro-structure
can be achieved and propose a strategy of gecko advance.
Based on the experimental scaling relations, we conjec-
ture that the same adhesion mechanism is operative at
every scale of the hierarchy and propose a simple model
justifying the observed power law force-length relations.
II. DISCRETE MODEL
To describe an elemental fibrillar adhesion layer we use
the minimal Bishop-Peyrard (BP) model [14 -- 16]. Con-
sider a chain with n + 1 particles interacting through n
linear springs and bound to a rigid substrate by break-
able elastic elements (see Fig.1). Assume that the par-
ticles can move only in the vertical direction and denote
by ui the displacement of the particle i. The coupling
between individual adhesive devices is controlled by the
elastic energy
φG(δi) =
1
2
Gδ2
i ,
where G is the (shear) stiffness, δi = (ui+1 − ui)/l is the
strain and l is the spring length. The cohesive energy at
the micro scale can be modeled by a piece-wise quadratic
function
(cid:40) 1
2 ku2
i ,
γb,
φk(ui) =
if ui < ur,
if ui ≥ ur,
where k is the extensional stiffness, γ is the adhesion
energy density, b is the out of plane spatial scale and
ur =(cid:112)2γb/k is the limit displacement.
The corresponding energy landscape is complex [16],
however, we are interested only in a set of local mini-
mizers that can be parametrized by the position of the
decohesion front ξ ∈ [0, n],
(1)
d − (i − 1)l
f
G
,
ui =
cosh[(n+3/2−i)η]l
2 sinh[(n+1−ξ)η] sinh[η/2]
2
i = 1, ..., ξ,
f
G
, i = ξ + 1, ..., n + 1.
(3)
Here the force is implicitly given by
f =
(2ξ − 1 + coth η
where η is a solution of
2nν2
2 coth[(n + 1 − ξ)η])
d
L
k
(4)
and
1 + l2/(2ν2) = cosh[η]
ν =(cid:112)
G/k
FIG. 1: Bishop-Peyrard model of decohesion.
We assume that the organism applies to the pad a lo-
calized loading. Denote by d the assigned displacement
at the point n = 1 and assume that all other points are
unloaded. To find the equilibrium force-displacement re-
lation f (d) we minimize the energy
n+1(cid:88)
n(cid:88)
Φ = l(
φk(ui) +
φG(δi)).
(2)
is the internal length scale, defining the size of the cohe-
sive zone.
i=1
i=1
FIG. 2: Equilibrium force-displacement curves for a BP system with n = 10 breakable links, L = 3, ur = 1, k = 1, and G = 1.
Solid lines indicate metastable branches. Bold lines indicate force-displacement paths associated with the overdamped limit
(a) and with global energy minimization (b).
In Fig.2 we show the metastable branches f (d; ξ)
parametrized by ξ; each branch ends at ¯d(ξ) satisfy-
ing uξ+1( ¯d(ξ); ξ) = ur. If the dynamics is overdamped
and the driving d(t) is quasistatic we obtain the loading-
unloading hysteresis indicated in Fig.2(a) by bold lines.
An alternative, hysteresis-free path of global energy mini-
mization implied in Kendall's model is shown in Fig.2(b).
Observe that in both cases the cohesion force exhibits a
characteristic plateau with which we can associate the
maximal force fm. The dependence of this threshold
on the stiffness of the pad G plays the main role in
the proposed mechanism. The quasistatic assumption,
which this model shares with Kendall's model, is sup-
ported by experimental observations that the attach-
ment/detachment rates are independent of the geckos
speed [17].
III. CONTINUUM LIMIT
A simple analytical expression for the function fm(G)
can be obtained in the continuum limit, which may be a
poor approximation at the scale of the whole foot (n ∼
5), but turns out to be fully adequate at the level of
n+1fd10.00.20.40.60.80.00.20.40.60.81.00.0 0.5 1.0 1.5 2.0 2.5 3.00.0 0.5 1.0 1.5 2.0 2.5 3.01.03
for L (cid:28) ν the cohesive zone spreads along the whole pad
and this regime can be associated with a pull off mode.
Interestingly, for L (cid:29) ν the area under the hysteresis
loop
Q = 2γbL
representing the detachment energy is exactly twice as
big as in the case when L (cid:28) ν; high dissipation at small
force is due to much larger displacement. In this sense
transition from pull off to peeling is similar to the tran-
sition from brittle to ductile fracture.
Our main assumption in what follows is that the stiff-
ness of the linear springs, mimicking the stiffness of the
gecko's pad, can be actively varied by the organism. The
feasibility of such control is clear from the fact that the
gecko rolls in for attachment (active shortening and thick-
ening of the digits) and rolls out for detachment (active
lengthening and thinning of the digits). It is also known
that digital hyperextension anticipates each attaching
and detaching event and that the musculo-tendinous sys-
tem may influence single lamellae in the process of con-
trolled rolling [13, 17]. Experiments with geckos [13],
frogs [19], and ants [20] also show that at the macroscale
the easy release is achieved through the localization of
the cohesive region. Yet another argument in support
of the active muscular control comes from the observa-
tion [21] that to simplify horizontal walking geckos keep
the hyperextended state and start activating the attach-
ment mechanism only at sufficiently high slope requiring
stronger adhesion.
To get a rough estimate of the required stiffness vari-
ation we observe that the adhesion force in geckoes is
about one order of magnitude less than the detachment
threshold [22]. According to (7) this corresponds to two
orders of magnitude in stiffness variation, which is com-
patible with the data on geckos forced to detach [23].
Different physical mechanisms may be employed to regu-
late the coupling among adhesive fibrils at different struc-
tural levels. Thus, at the cellular level the dynamic fil-
amentous actin network is known to be very soft at low
stress, but can stiffen up to three orders of magnitude in
response to stresses [24 -- 26]. Such stresses can be gen-
erated internally through molecular motors; moreover,
constant remodeling allows the cytoskeleton to remain
in a marginally stable state and easily switch between
softening and stiffening [27, 28]. At elevated stresses a
quick transition to softening may also be related to the
unfolding of the cross-linkers such as filamin [29]. Within
muscle sarcomeres the effective stiffness can vary with the
number of myosins attached to actin fibers and will also
be affected by the unfolding-refolding transition in titin
[30]. Notice also that parameter ν controlling the mode
of detachment can vary not only because of the stiffness
G but also because of the stiffness k. The latter depends
on the aspect ratio of the adhesive elements and may be
controlled by the organism through capillarity induced
self-assembly, modulated by secretion or evaporation of
liquids responsible for the interaction between the fibrils
0
(cid:90) L
d − x
cosh L−x
cosh L−ζL
ν
FIG. 3:
Stiffness dependence of the adhesion threshold
fm(G) for a system with ur = 1 mm, k = 1 MPa and
L = 10 mm. Insets: crack configurations illustrating local-
ized (L/ν = 100) and diffuse (L/ν = 1) cohesive zones.
the setae (n ∼ 50). To study this limit we fix the total
length L = nl and assume that n → ∞ and l → 0. The
continuum energy takes the form
Φ(u) =
(φk(u) + φG(u
(cid:48))) dx
(5)
and its minimization is straightforward [16, 18]. Given
the loading u(0) = d we obtain
uζ(x) =
ζL (d − ur)
ur
ν
if x ∈ (0, ζL),
if x ∈ (ζL, L),
(6)
where ζ is the detached fraction of the pad which
parametrizes the relation between the boundary displace-
ment
d = ur + ζL(f /G)
and the applied force
f =(cid:112)2γ b G tanh[L(1 − ζ)/ν].
fm(G) =(cid:112)2γ b G tanh(L√k/√G).
The maximum value of the force fm, representing the
detachment threshold, is attained at ζ = 1 and d = ur:
(7)
In the limit when the external length scale is much larger
than the internal length scale (L (cid:29) ν) we get the asymp-
totics
In the opposite case we obtain
(cid:112)2γ b G.
(cid:112)2γ b kL.
fm ∼
fm ∼
The value of the critical force and the structure of the
associated cohesive layers is shown in Fig.3. Notice that
for L (cid:29) ν the crack has a narrow tip and this regime can
be associated with the peeling mode of fracture. Instead
0510152000,10,20,30,40,50,60,70,80,91.0kpL≫νfmaxL≪ν(N)00.511.522.530.0750.080.0850.090.0950.10.1050.110.115ν0u0.1L / =1pull off0 1 2 3x00.511.522.5300.511.522.53105G(N)f mL / =100x0 5 10 15 νpeeling0.01120u0 2 3[31, 32].
At the macroscale, the variability of shear modulus
may be associated with the reversible development of mi-
cro defects inside the tissue architecture with subsequent
internal healing of this damage through remodeling. Fol-
lowing a classical approach in damage mechanics [33, 34],
we can introduce an internal variable α, with α = 0 repre-
senting the undamaged state (stiffest configuration) and
α = 1 -- the damage-saturated state (most compliant con-
figuration). Then G = G(α) with G(cid:48)(α) < 0. Since the
detachment force depends on the level of damage
fm(α) = fm( G(α)),
while the critical displacement does not (dm = ur), we
can write the effective elastic stiffness for the attached
state in the form
(cid:113)
k G(α) tanh[L√k/ G(α)].
E(α) =
Observe that E(cid:48)(α) < 0.
Schematic representation of
FIG. 4:
the proposed
attachment-detachment strategy (see text). The inset shows a
generic force-displacement path associated with stiffness vari-
ation AB and the decomposition of the corresponding external
work W into the elastic energy Φ and the dissipated energy
Q.
IV. ATTACHMNT-DETACHMENT CYCLE
We can now propose a mechanical strategy of geckos
advance. Assume that only one scale of the fibrillar mi-
crostructure is involved, that the response is quasistatic
and overdamped, and that the continuum limit is valid.
To fix ideas, consider ceiling walking and suppose that
the cycle of attachment-detachment starts at an attached
state A (see inset in Fig. 4) where the pad is in the
stiff configuration G(0) associated with the high thresh-
old fm(0). To attain the detachment below this thresh-
old, gecko can induce a reversible "damage process" de-
scribed by a force-displacement relation f = ¯f (d) where
d = d(α) and E(α) = ¯f ( d(α))/ d(α). In the process of
stiffness variation the external work
W =
¯f ( d) d d > 0
4
(cid:90) d(α)
P/ E(0)
(cid:90) α
(done by gravity) is partially stored as elastic energy
[Φ] = P 2/(2 E(0)) − E2(α) d(α)/2
and the rest is dissipated into heat
Q = −
0
(cid:48)(α) d 2(α)dα.
E
1
2
To minimize dissipation while maintaining a stable me-
chanical response with ¯f(cid:48)(d) ≥ 0, the animal must ensure
that damage advances at constant force ¯f (d) = P (AB in
Fig.4). In this case exactly half of the work is dissipated
and we can write
[Φ] = Q = (P/2)(ur − P/ E(0)) > 0.
Damage at constant stress has been observed in many
rubber-like materials and linked to inherent energy non-
convexity [35]; reversible structural changes at constant
stress are also characteristic for muscle tetanus [36].
After the critical force is decreased, the pad can be
pulled away by peeling (path BC in Fig.4). The detach-
ment ends with an abrupt decohesion at point C in Fig.4.
In order to reattach, the gecko can follow a reverse path
EF. As the foot is placed on the surface, the displacement
d is gradually decreased and the attachment takes place
at the point F through "inverse peeling" at d = ur. An
instantaneous force jump brings the system to the point
G and allows the animal to place some weight on the
foot. To secure a robust attachment, the gecko must re-
verse the damage and induce active stiffening (trajectory
GH).
The system heals the damage by remodeling the dam-
aged configuration with α = 1 back into the virgin con-
figuration with α = 0 while increasing stiffness and de-
creasing displacement. This requires work which is now
done by the gecko. The energy comes from metabolic
sources M < 0 and is released due to elastic unloading
[Φ] = ( fm(1)/2)( fm(1)/ E(0) − ur) < 0.
If we again assume that healing (remodeling) takes place
at constant stress (minimum metabolic energy path),
we obtain [Φ] = M . Experiments show that geckos in
the compliant state have a very low detaching thresh-
old fm(1) [37] which means that the metabolic energy
required for such stiffness increase is also low.
After the state of high stiffness is reached (point H in
Fig.4) the peeling mode is deactivated because the force
required for detachment fm(0) is now large. Therefore
more weight can be shifted to this foot (path HI) and
another foot can undergo the detaching-attaching cycle.
Overall, the detachment process BCD with decreasing
force on the two detaching feet must take place simul-
taneously with the attachment process FGHA involving
the other two reattaching feet [17].
00.511.522.533.50.40.50.60.70.80.911.11.2fd^^^^^^(α)(α)(α)m05d(mm)CDEB(N)pull offrBpeelingmHGrIdΦ( )^_Q30A(1)FAf( )Eu(0)EuEdP(1)E(0)fαff d25
i+1 = nil2
which accommodates fractal roughness. If at hierarchi-
cal level i, with spatial scale li, the system is marginally
stable against buckling, then fi = cEl2
i , where E is the
elastic modulus and the constant c depends on the shape
of the cross section, aspect ratio, and the boundary con-
straints. We assume the simplest allometric law when
both c and E are scale invariant. Now, if the fibrils at
a finer level i cover the tips of the fibrils at the coarser
level i + 1 densely, which is known as Leonardo's rule
[42], then l2
i where ni is the number of fibrils
at the level i (see the scheme at the bottom right corner
of Fig. 5). Since fi+1 = nifi, we obtain fi+1 = cEl2
i+1,
which means that the force is critical also at the i+1 level.
In view of our neglect of collective modes of instability
[43], we can only tentatively conclude from this reason-
ing that the observed scaling supports the idea that the
whole structure can be marginalized simultaneously. An
important feature of this scaling, however, is that stress
is uniform which has been proposed previously as a cri-
terion of optimality in several biological and engineering
systems [44].
The power law scaling is indicative of a scale-free de-
tachment mechanism. We can model it in our framework
by the appropriate renormalization of the parameters in
a scale-generic BP model shown in the upper corner of
Fig.5, where ue is the elastic threshold and ur is the
detachment displacement. The parameters ue, ur, and
k can be computed at each scale iteratively by using a
series of nested BP models, whereas the parameter G
characterizing the elastic coupling can be chosen at each
scale to match the experimentally measured detachment
threshold.
At the smallest scale of a spatula the adhesive proper-
ties of the fibrils can be described by the energy density
(1) with parameters ur and k available from experiment
[6]. The behavior at the next scale (setae) can be ob-
tained numerically by simulating an overdamped gradi-
ent flow dynamics for a quasistatically driven BP system
with n = 50 spatulae; the value of G is chosen to ensure
the maximum measured adhesion force of 40µN [7]. The
overall behavior at this scale matches the experimental
results in [7] showing an elastoplastic range ending with
an abrupt detachment. At the level of a lamella, we con-
sider n = 50 elastic-plastic elements with constitutive
parameters obtained from the spatulae scale simulations.
At the next level of a toe, we need to model lamellae
with n = 20 and finally at the level of a foot n = 5 (toes)
and the problem becomes strongly discrete. The results
at this last level match observations showing digitized
detachment of the toes [13].
VI. CONCLUSIONS
By studying a prototypical system, we have shown that
both the force threshold and the dissipation associated
with reversible adhesion can be modified by active con-
trol of the coupling among individual adhering elements.
FIG. 5: Scale free nature of the adhesion mechanism. Solid
line interpolates the experimental values of the adhesion forces
[1, 3, 6 -- 8, 10, 23, 37, 39]; dashed line, of the friction forces
[37]. Insets show nested computations based on the BP model
with adhesion links behaving as shown in the right upper
corner. At the smallest scale of the elemental fibrilla ur = 200
nm, L = 103 nm and k = 100MPa. The parameters ue,
ur, and k at larger scales have been computed iteratively.
Other parameters: fibrilla → seta, n = 50, L = 100 nm,
G = 0.2 mN; seta → lamella, n = 50, L = 0.5 mm, G = 1 N,
lamella → pad, n = 20, L = 3 mm, G = 300 N, pad → foot:
n = 5, L = 20 mm, G = 20 N.
V. HIERARCHICAL ARCHITECTURE
To understand the role played by the hierarchical
micro-fibrillar architecture of the adhering pad [10, 38]
we compare adhesion forces at different spatial scales
(see Fig.5). Using the experimental data [1, 3, 6 --
8, 10, 23, 37, 39] we may deduce a scaling relation
fm ∼ lβ,
with exponent β = 2 ± 0.5. In Fig.5 we show the power
law together with the scaling of friction forces reported
in [37]; our results are consistent with the observation
that the friction force is usually five times higher than
the adhesion force [23].
While the exact origin of these empirical power law re-
lations is not known, we can propose the following plausi-
ble explanation. Following [40] we suppose that the archi-
tecture of the fibrillar adhesive system is designed to en-
sure that in contact with rough surfaces the fibrils buckle
simultaneously at all scales to ensure maximum folding
fmln_l0md2*1044*1042*1094*109ffd2*105*10610*106f34*103dll1002002010dfru102*10i_i+uef102*104*107710102*104104dflnl01d102*1033ffPRIN 2010-11: "Dinamica, stabilit`a e controllo di strut-
ture flessibili".
6
The possibility of the ensuing multi-path adhesion [45]
reconciles strong binding with easy debinding, which is
at the base of the observed agility of lizards and insects
running on inclined surfaces. The proposed mechanism
has a scale-free hierarchical structure, which is typical for
biological systems at all levels of organization.
Acnowledgments
The authors thank F. Maddalena and D. Percivale for
helpful discussions. G.P. work has been supported by
[1] W. Sun, P. Neuzil, T. S. Kustandi, S. Oh, and V. D.
intosh. Science, 315, 370 (2007).
Sam- per. Biophysical J., 89, L141 (2005).
[2] Y. Tian, N. Pesika, H. Zeng, K. Rosenberg, B. Zhao,
P. McGuiggan, K. Autumn, and J. Israelachvili. PNAS,
103, 19320 (2006).
[3] B. Chen, P. D. Wu, H. Gao. Proc. Roy. Soc. Lon., 464,
1639 (2008).
[4] B. N. J. Persson. MRS Bulletin, 32, 486 (2007).
[5] N. M. Pugno, F. Bosia and T. Abdalrahman. Phys. Rev.
E, 85 011903 (2012).
[6] G. Huber, S. N. Gorb, R. Spolenak and E. Arzt. Biol.
Lett., 1, 2 (2005).
[7] K. Autumn, Y. A. Liang, S. T. Hsieh, W. Zesch, W. P.
Chan, T. W. Kenny, R. Fearing and R. J. Full. Nature,
405, 681 (2000).
[25] G. H. Koenderink. PNAS, 106, 15192 (2009).
[26] K. E. Kasza, G. H. Koenderink, Y. C. Lin, C. P. Broed-
ersz, W. Messner, F. Nakamura, T. P. Stossel, F. C.
MacKintosh, and D. A. Weitz. Phys. Rev. E, 79 041928
(2009).
[27] J. C. Martens, M. Radmacher. Pflugers Arch., 456, 95
(2008).
[28] E. M. Reichl,Y. Ren,M. K. Morphew,M. Delannoy,J. C.
Effler,K. D. Girard,S. Divi,P. A. Iglesias,S. C. KuoandD.
N. Robinson. Curr. Biol., 18, 471 (2008).
[29] B. A. Di Donna, A. J. Levine. Phys. Rev. E., 75041909
(2007).
[30] W. A. Linke, M. Kruger. Physiology, 25, 186 (2010).
[31] T. Eisner, D. J. Aneshansley. Proc. Natl. Acad. Sci., 97,
[8] K. Autumn, A. Dittmore, D. Santos, M. Spenko and M.
6568 (2000).
Cutkosky. J. Exp. Biol., 209, 3569 (2006).
[32] B. Pokroy, S. H. Kang, L. Mahadevan, J. Aizenberg. Sci-
[9] A. K. Geim, S. V. Dubonos, I. V. Grigorieva, K. S.
Novoselov, A. A. Zhukov and S. Y. Shapoval. Nature
Mat., 2, 461 (2003).
ence, 323, 237 (2009).
[33] D. D'ambrosio, D. De Tommasi, D. Ferri, G. Puglisi. Int.
J. Engi. Sci., 46, 293 (2008).
[10] H. Gao, X. Wang, H. Yao, S. Gorb, E. Arzt. Mech.
[34] A. DeSimone, J. Marigo, L. Teresi. Eur. J. Mech.
Mater., 37, 275 (2005).
A/Solids, 20, 873 (2001).
[11] H. Yao, G. Della Rocca, P. R. Guduru, and H. Gao. J.
[35] D. De Tommasi, G. Puglisi, G. Saccomandi. Phys. Rev.
R. Soc. Interf., 5, 723 (2008).
Let., 100, 085502 (2008).
[12] B. Chen, P. D. Wu and H. Gao. J. Roy. Soc. Interf., 6,
529 (2009); Q. H. Cheng, B. Chen, H. J. Gao and Y. W.
Zhang. J. Roy. Soc. Interf. 9, 283 (2012).
[13] A. P. Russell. Integr. Comp. Biol , 42, 1154 (2002).
[14] M. Peyrard. Nonlinearity, 17, R1 (2004).
[15] N. Theodorakopoulos, M. Peyrard, and R. S. MacKay.
Phys. Rev. Lett. 93, 258101 (2004).
[36] A. M. Gordon, A. F. Huxley, F. J. Julian. J. Physiology,
184, 170 (1966).
[37] K. Autumn. Properties, Principles, and Parameters of
the Gecko Adhesive System. In Biological adhesive,
edited by Smith, A. and J. Callow (Springer-Verlag,
Berlin 2006).
[38] T. Kim, B. Bhushan. J. Adhes. Sci. Technol., 21, 1
[16] F. Maddalena, D. Percivale, G. Puglisi and L. Truski-
(2007).
novsky Cont. Mech. Therm., 21, 251 (2009).
[17] K. Autumn, S. T. Hsieh2, D. M. Dudek, J. Chen, C.
Chitaphan, and R. J. Full. J. Exp. Biol., 209, 260 (2006).
[18] J. J. Marigo, L. Truskinovsky Cont. Mech. Therm., 16,
391 (2004).
[19] G. Hanna, W. Barnes. J. Exp. Biol., 155, 103 (1991).
[20] W. Federle, T. Endlein. Antropod Struct. & Develop., 33,
67 (2004).
[21] A. P. Russell and T. E. Highman. Proc. Roy. Soc. B,
276, 3705 (2009).
[22] Z. Dai, Z. Wang and A. Ji. J. Exp. Biol., 214, 703 (2011).
[23] N. Pugno, E. Lepore, S. Toscano and F. Pugno. J. Ad-
hesion, 87, 1059 (2011).
[39] D. J. Irschick, J. B. Losos, Biol. J. Lin. Soc., 59,
21 (1996); M. Nosonovsky and B. Bhushan. Multi-
scale Dissipative Mechanisms and Hierarchical Surfaces.
NanoScience and Technology,, (Springer, New York,
2008), pp. 231-242; B; B. Zhao, N. Pesika, K. Rosen-
berg, Y. Tian, H. Zeng , P. McGuiggan, K. Autumn, and
J. Israelachvili. Langmuir, 24, 1517 (2008); K. Autumn,
C. Majidi, R. E. Groff, A. Dittmore and R. Fearing. J.
Exp. Biol., 209, 3558 (2006).
[40] A. Jagota, S. J. Bennison.
Integr. Comp. Biol., 42,
1140 (2002); B.N.J. Persson. J. Chem. Phys., 118, 7614
(2003); N. J. Glassmaker, A. Jagota, C. Y. Hui, J. Kim.
J. R. Soc. Interface, 1, 23 (2004).
[24] D. Mizuno, C. Tardin, C. F. Schmidt, and F. C. MacK-
[41] T. A. McMahon, R. E. Kronauer J.Theor. Biol., 59, 443
(1976).
[42] C. Eloy. Phys. Rev. lett., 107, 2581101 (2011).
[43] M.Rodriguez, E. de Langre, B. Moulia Am. J. Botany.,
95, 1523 (2008).
93 (2012); G. A. Holzapfel, R. W. Ogden. Proc. R. Soc.
A, 466, 1551 (2010); M. P. Bendsoe, O. Sigmund. Topol-
ogy optimization, (Springer-Verlag, Berlin, 2003).
[45] D. Bartolo, I. Derenyi, A.Ajdari, Phys. Rev. E, 65,
[44] T. A. McMahon, Muscles, reflexes and locomotion,
Princeton, 1984; M. Destrade et al. J. Theor. Biol., 303,
051910 (2002).
7
|
1304.1747 | 1 | 1304 | 2013-04-05T15:35:36 | Condensation of circular DNA | [
"physics.bio-ph",
"cond-mat.soft",
"physics.chem-ph"
] | A simple model of a circularly closed dsDNA in a poor solvent is considered as an example of a semi-flexible polymer with self-attraction. To find the ground states, the conformational energy is computed as a sum of the bending and torsional elastic components and the effective self-attraction energy. The model includes a relative orientation or sequence dependence of the effective attraction forces between different pieces of the polymer chain. Two series of conformations are analysed: a multicovered circle (a toroid) and a multifold two-headed racquet. The results are presented as a diagram of state. It is suggested that the stability of particular conformations may be controlled by proper adjustment of the primary structure. Application of the model to other semi-flexible polymers is considered. | physics.bio-ph | physics | Condensation of circular DNA
E.L.Starostin1
Department of Civil, Environmental & Geomatic Engineering,
University College London, Gower Street, London WC1E 6BT,
UKa)
(Dated: 13 November 2018)
A simple model of a circularly closed dsDNA in a poor solvent is considered as an
example of a semi-flexible polymer with self-attraction. To find the ground states,
the conformational energy is computed as a sum of the bending and torsional elastic
components and the effective self-attraction energy. The model includes a relative
orientation or sequence dependence of the effective attraction forces between different
pieces of the polymer chain. Two series of conformations are analysed: a multicovered
circle (a toroid) and a multifold two-headed racquet. The results are presented as a
diagram of state. It is suggested that the stability of particular conformations may be
controlled by proper adjustment of the primary structure. Application of the model
to other semi-flexible polymers is considered.
PACS numbers: 87.15.bk, 36.20.-r, 87.15.A-
3
1
0
2
r
p
A
5
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
4
7
1
.
4
0
3
1
:
v
i
X
r
a
a)Electronic mail: [email protected]
1
I.
INTRODUCTION
When a double-stranded DNA (dsDNA) finds itself in a solution with a condensing agent,
it tends to take on a conformation that would minimise its exposure to the environment and,
correspondingly, maximise the sites of self-contact1,2. Since dsDNA is a rather stiff molecule,
variation of its shape implies a certain energy cost which is usually described as the elastic
energy of bending and twisting. The aim of this paper is to derive simple estimates that help
us to decide what conformation is preferable for a circularly closed dsDNA in terms of its
total energy. The molecule is assumed to be unknotted and torsionally relaxed when in pure
water. We are interested in finding stable ground states and do not consider temperature
effects.
If we assign orientation to the centreline of the dsDNA, then two ways of condensation
are logically possible: the remote intervals of the dsDNA chain can align next to each other
such that their orientation is either the same or opposite. The first case is best achieved in a
double- (or multiple-)covered circle, while the second realises as a straight line interval with
two loops at the ends (Fig. 1). Thus we come to familiar toroidal and racquet-like shapes3,4.
Here we consider the series of the simplest possible conformations: a toroid and a two-headed
racquet. The latter does not necessarily imply a significant writhe, while the first has its
writhing number growing linearly with its number of coils5. Thus, the writhing number of
a toroid with N coils equals ±(N − 1), if we neglect the thickness of the polymer, while a
multiple-covered two-headed racquet may be made to have its writhing number vanish. We
assume that the linking number is conserved, therefore, the non-zero writhe is compensated
by twist which should be accounted for in the calculation of the elastic energy. Recall that,
to avoid accumulation of twist in long ropes, sailors store them in a figure-8 fashion, a similar
trick is used when mountaineers "butterfly" the coils of their climbing ropes. These coilings
are analogous to the two-headed racquets.
Change of the relative orientation of contacting pieces of dsDNA is only one, most pro-
nounced, feature of two types of condensate. Actually, the coupling of the particular sites
that interact with each other differ as well. On the other hand, the interaction potentials of
two homologous and nonhomologous DNA duplexes are found to be significantly different6.
This suggests that the monomer sequence may be intentionally tuned to amplify the effective
attraction forces for one specific conformation. Toroids and racquets may be considered as
2
analogs of the secondary structures of single-stranded nucleic acids where dsDNA plays part
of a single stranded polymer and the hydrogen bonds between complementary nucleotides
are replaced with the electrostatic attraction. The secondary structure of RNA or ssDNA
may be controlled by varying their primary structure. Similarly, the primary structure may
affect the relative stability of toroids and racquets as the dsDNA condensates. Note that the
sequence-dependent DNA condensation was studied on double-stranded poly(dG-dC)·(dG-
dC) (GC-DNA) and ds poly(dA-dT)·(dA-dT) (AT-DNA)7. The effects due to the relative
orientation of the contacting pieces or to their specific sequence contents do not exist for
such homogeneous primary structures.
As known from earlier work (both theoretical and experimental) for open polymer
chains3,4,7 -- 9, the racquets can only be so-called metastable structures which means that
they always lose to toroids though their conformational energy may not differ much from
that of toroids. Numerical simulation suggests that interaction with adsorbing surface tend
to stabilise metastable racquet-like structures10. Experiments with condensation of circular
plasmids (closed un/nicked and open) also show intermediate racquet-like shapes which
seem to occur more often for unnicked DNA11. What if the racquets be given the upper
hand by building a DNA with specially tuned mirrored homologous subsequences in order
to increase their interaction potential? The competing toroids must not feel it which could
cause loss of their role as most stable structures. The model considered in this paper ex-
plores such a possibility. To simplify analysis it will be assumed that all the contacting sites
in racquets have uniform interaction potential which can differ from that of toroids where
the interaction forces are also assumed to be uniform.
It was recently demonstrated by
numerical simulation that presence of a direction-dependent potential affects condensation
of semi-flexible polymers12.
II. CONFORMATIONAL ENERGIES
We use a model of the semi-flexible polymer chain that includes two parts of the confor-
mational energy: the elastic energy and the self-interaction energy4. The balance between
these two components determines the geometry of the ground-state configuration.
For a polymer molecule of length L, we describe the bending energy as that of a slender
3
d
e
f
g
a
b
c
FIG. 1. Two ways of condesation of a circular dsDNA: N -covered circles (toroids) (a: an initial
conformation with N = 1, b: N = 2, c: N = 3) and M -folded two-headed racquets (d: M = 1,
e,f,g: M = 2). Red colour marks parallel self-alignment, green antiparallel. The shapes (e), (f)
and (g) differ with the writhing number, e: Wr = ±1, f: Wr = ±1 ± 1, g: Wr = ±1 (unknotted).
elastic rod:
Ubend =
B
2
L
Z
0
κ2(s) ds,
(1)
where B is the bending stiffness and κ(s) is the curvature of the centreline. Similarly, the
torsional energy is computed as
ω2(s) ds,
(2)
Utors =
C
2
L
Z
0
4
where C is the torsional stiffness and ω(s) is the twist rate of the polymer. We do not
account for dependence of the elastic moduli on the variations of the sequence of monomers.
Neither stretching of DNA is included in the model as of minor importance.
Each piece of the polymer chain may be in contact with up to six other pieces. Conden-
sation is possible when every such contact decreases the conformational energy. In other
words, the polymer prefers to contact with itself in order to screen its surface from ex-
position to the solvent. To account for the self-attraction, we employ the concept of the
coordination mumber αN . The doubled coordination number 2αN counts the number of
possible binding sites that are left non-occupied by N parallel polymer strands which fill in
the triangular lattice in the orthogonal cross section4. There exists an exact formula for the
minimal coordination number13
αN = l√12N − 3m ,
where ⌈·⌉ stands for the ceiling function: for x ∈ R, ⌈x⌉ is defined as the smallest integer
greater than or equal to x. In particular, α1 = 3, α2 = 5, α3 = 6, and so on. Let γ be the
energy gain of one contact per length. We shall assume that γ > 0. Then for an N-fold
parallel bundle of length LN the interaction energy is proportional to the difference of the
sum of the coordination numbers of N separate chains and the coordination number of them
packed together
UN = −γ(α1N − αN )LN .
(3)
In what follows, all the lengths and energies will be presented in dimensionless form by
normalising them on the condensation length Lc = pB/γ and the condensation energy
Uc = √Bγ, respectively4. We shall denote the normalised contour length of the polymer
chain by λ = L/Lc. The analysis is carried out under a simplified assumption of the zero
thickness of the polymer.
A. Multi-covered circle (toroid)
In this work we consider only circularly closed polymers. The reference conformation
is the single-covered circle which minimises the elastic energy. DNA is torsionally relaxed
so that the linking number of its strands Lk = Lk0 = T w0, where T w0 corresponds to the
5
intrinsic twist of the B-form. The normalised bending energy is ubend = 2π2/λ. There is no
self-interaction and, correspondingly, Eq. (3) gives us zero for N = 1.
The circularly closed DNA may form a N-covered circle or a toroid without being nicked.
The size of the toroid can only take discrete values contrary to the open case. Let ρ be the
radius of the circle which is traced N times by the DNA before it closes onto itself so that
L = 2πρN. Then the curvature κ = 1/ρ = const and the bending energy Eq. (1) equals
Ubend = B
2
1
ρ2 L = 2π2B N 2
L .
Since the linking number is conserved, we may estimate the change of twist as T w −
T w0 = −Wr , where the writhing number of the N-covered circle can be estimated as
±(N − 1). We further assume that the twist is uniformly distributed along the length of the
polymer, i.e. ω(s) = ±2π(N − 1)/L. Then the torsional energy can be found from Eq. (2):
Utors = C
, where
, and after nomalisation we have utors = 2π2c (N−1)2
2 ω2(s)L = 2π2C (N−1)2
L
λ
we introduced c = C/B. Note that in order to model a nicked DNA one can formally set
c = 0 to kill the torsional energy.
The fact that the polymer touches itself uniformly along all its length allows us to apply
Eq. (3) for LN = L/N. Then the normalised interaction energy is simply uint = (αN −
Nα1)λ/N. Summing up all three energies gives the total
ut = ubend + utors + uint =
=
2π2
λ
(N 2 + c(N − 1)2) +(cid:16)αN
N − 3(cid:17) λ.
(4)
The above expression coincides with Eq. (4) in Ref.14 for c = 0.
B. Two-headed racquet
To minimise the bending, the heads should be identical. Then the elastic energy reduces
to the doubled bending energy of a racquet head which is assumed to be a bundle of M
identical plane inflexional elasticae15. Let the contour length of the head be denoted by χ
which is the only parameter the normalised bending energy of a single component depends
on: uhead = a/χ, where a = 4ξ2(k0)(2k2
the elliptic integral of the first kind of modulus k
0−1), ξ(k0) = 2K(k0)−F(cid:16) 1
k0√2
, k0(cid:17)4. Here F denotes
F(z, k) =
z
Z
0
1
√1 − t2√1 − k2t2
dt,
6
and K(k) ≡ F(1, k) is the complete integral. The elliptic modulus k0 is found as a root of
the equation 2η(k0) = ξ(k0), where η(k) = E(sn(ξ(k), k), k) = 2E(k) − E(cid:16) 1
, k(cid:17) and E is
k√2
the elliptic integral of the second kind of modulus k
E(z, k) =
z
Z
0
√1 − k2t2
√1 − t2
dt,
E(k) ≡ E(1, k). Numerical solution gives k0 ≈ 0.8551 and a ≈ 18.3331.
To compute the interaction energy of the head, we apply Eq. (3) which becomes uM =
(αM − Mα1)χ. The straight handle is a 2M-bundle of normalised length λ/(2M) − χ and
Eq. (3) yields u2M = (α2M − 2Mα1)(λ/(2M) − χ). The total energy is the sum 2Muhead +
uM + u2M . Like for an open polymer, the size of the racquet head may take continuous
values. For given length, the optimal length of the head equals χ = q 2aM
αM −α2M +3M and
finally we have u = 2p2aM(αM − α2M + 3M) + (α2M /(2M) − 3)λ. Note that the optimal
racquet may only exist for polymers not shorter than 2Mχ. Its actual size is proportional,
for given M, to the condensation length (contrary to discrete diameters of the toroids made
of a closed polymer).
To account for the difference in the effective attraction forces for parallel and antipar-
allel alignments we distinguish the interaction energy coefficient for the racquet γr. It is
convenient to introduce a new parameter ν, ν2 = γr/γ which measures the orientation dif-
ference. ν2 > 1 corresponds to stronger attraction in the antiparallel configuration (as in
the racquet handle). On a triangluar lattice the antiparallel arrangement is frustrated and
in the ground state one third of all neighbouring pairs have the same orientation, similarly
to antiferromagnetic materials16. When computing the value of ν, this property should be
taken into consideration.
To keep the common normalisation, we rewrite the racquet energy as
ur = 2p2aM(αM − α2M + 3M)ν + (α2M /(2M) − 3)λν2.
(5)
III. COMPARISON OF CONFORMATIONS
Now we are able to compare the energy of the conformations as functions of the polymer
length and the orientation parameter ν2. We begin with N- and P -covered circles which
are members of a sequence of toroidal shapes. Comparison of the conformational energies
7
(Eq. (4)) leads to the critical length
λ2
N P = 2π2NP (N − P )
(N + P )(c + 1) − 2c
NαP − P αN
.
(6)
By use of this expression we can build a sequence of optimal toroids {Ni} with critical
lengths {λi}. For a polymer of length λ, the Ni-circle is optimal among all toroidal shapes
if λi−1 < λ < λi. We first set N0 = 0, λ0 = 0 and start with a circle with N1 = 1. Next,
assuming that we know the optimal sequence up to i = I (i.e. we know Ni, i = 0, . . . , I, and
λi, i = 0, . . . , I − 1) we compute λI = λNI P ⋆ = min
to check that ∂λut(λ, P ⋆) < ∂λut(λ, NI) for λ = λI and hence there exists ǫ > 0 such that
the inequality holds for λ ∈ [λI, λI + ǫ). Therefore the NI+1-toroid is optimal for λ > λI in
some right neighbourhood of λI. Now we can apply the same procedure to find λI+1, NI+2
λNI P , NI+1 = max
P ⋆. It is easy
P >NI
λNI P ⋆ =λI
and so on. The first members of the optimal toroid sequence are (Fig. 2):
St = {1, 2, 3, 4, 5,
7, 10, 12, 14, 16,
19, 24, 27, 30, 33,
37, 44, 48, 52, 56, . . .}.
(7)
The critical lengths are shown in Fig. 3 for four different values of the torsional to bending
stiffness ratio c (note that, as can be seen from Eq. (6), the optimal sequence St does not
depend on c for nonnegative values).
The sequence Eq. (7) may be described as a series of quintuples
St = {3n(n + 1) + 1,
{(3n + m + 1)(n + 1), m = 1, 2, 3, 4},
n = 0, 1, 2, . . .}.
As proposed in Ref.14, the sequence is made with the numbers that satisfy α(N) = α(N −
1) = α(N + 1)− 1. This characterisation must be corrected: numbers that can be expressed
as (n + 1)(3n + 1), n = 1, 2, 3, . . ., must be removed from the optimal sequence.
Next we compare the energy of the racquet-like shapes by using Eq. (5) for M and Q.
We come to the equation for the critical value of the effective length of the polymer µ ≡ νλ
8
FIG. 2. The sequence {Ni} for the optimal Ni-covered circles (toroids).
for the M- and Q-racquets
µM Q =
4√2aMQ
Mα2Q − Qα2M hpM(αM − α2M + 3M)−
−qQ(αQ − α2Q + 3Q)(cid:21) .
Similarly to the toroids we build a sequence of optimal racquets {Mj} with critical lengths
{µi}. For a polymer of length µ, the Mj-racquet is optimal among all racquet shapes if
µj−1 < µ < µj. To start the sequence we set M0 = 0, µ0 = 0, and M1 = 1 corresponds to
a single-covered two-headed racquet. Then, assuming that we know the optimal sequence
up to j = J (i.e. we know Mj, j = 0, . . . , J, and µj, j = 0, . . . , J − 1) we find µJ =
Q⋆. As seen from Eq. (5), the energy of an M-
µMJ Q⋆ = min
Q>MJ
µMJ Q, MJ+1 = max
µMI Q⋆ =µJ
9
FIG. 3. The sequence of critical lengths {λi} for the optimal Ni-covered circles (toroids) for four
different values of the relative torsional stiffness c.
racquet ur(µ, M), for fixed ν, is a linear function of µ. Moreover, it is an easy exercise
to prove that the function M(αM − α2M + 3M) monotonically increases with M so that
we have ur(0, Q) > ur(0, M) for Q > M. This means that if the graphs ur(µ, Q) and
ur(µ, M), Q > M, intersect each other at some positive µM Q, then ur(µ, Q) < ur(µ, M) for
µ > µM Q. Therefore the MJ+1-racquet is optimal for µ > µJ in some right neighbourhood
of µJ . Now we iterate the procedure to compute µJ+1, MJ+2 and so on. The first members
of the optimal racquet sequence are: 1, 2, 3, 4, 5, 6, 7, 8, 12, 15, 18, 20, 22, 24, 26, 28, 30, 35 . . .
(Fig. 4). The critical lengths are shown in Fig. 5.
The last comparison to make is between an N-covered circle and an M-racquet. Equating
10
FIG. 4. The sequence {Mj} for the optimal Mj-racquets.
energies taken from Eqs. (4) and (5) leads to the quadratic equation
(3 − α2M /(2M))µ2 − 2p2aM(αM − α2M + 3M)µ−
− (3 − αN /N)λ2 + 2π2(N 2 + c(N − 1)2) = 0.
(8)
Consider first the case N = 1. The coefficient of the λ2 term vanishes and we have a quadratic
equation solely for µ.
It has two positive roots but only the greater one is important:
µ1M = √2(paM(αM − α2M + 3M) +paM(αM − α2M + 3M) − π2(3 − α2M /(2M)))/(3 −
Clearly, we are most interested in µ11 = 2(√2a + √2a − π2) ≈ 22.4636, because for M ≥ 2
α2M /(2M)). Thus, the M-racquets win over simple circles if their length exceeds µ1M .
the simple circle is not competitive and the racquets should be compared with themselves
11
FIG. 5. The sequence of critical lengths {µj} for the optimal Mj-racquets.
and multicovered circles. In the general case N ≥ 2, Eq. (8) describes a hyperbola and it
is convenient to resolve it for λ as a function of µ. If the actual λ exceeds this value, the
energy of the M-racquet is higher than of the N-circle.
Our aim is to represent the results of energy comparison as a diagram of state in plane
(µ, λ). We begin with relatively short polymers (Fig. 6). A bisector divides the first quadrant
of the parameter plane into two triangular domains. The upper one, where λ exceeds µ,
corresponds to ν2 < 1, i.e. the stronger attraction for toroidal arrangements of DNA. Clearly,
the racquets have no chance to compete with toroids under such a condition unfavourable
for them. As the polymer is getting longer, the optimal condensate has a growing number of
circle coverings. The bisector itself corresponds to the equal strength of attraction potential
12
irrespective of the DNA primary structure. When the racquet-like shapes are benefited with
ν > 1, they appear on the surface. In any case their length should exceed the minimum
µ11. With growth of the polymer chain, the optimal racquets acquire more dsDNA strands.
Similarly, toroids increase their number of coils as λ becomes larger. In the (µ, λ) plane
the domain of the optimal Ni-toroid is bordered by two parallel straight lines λ = λi−1 and
λ = λi with the neighbouring Ni−1- and Ni+1-toroids. It also has a border with one or more
racquet domains. Each piece of this border is a part of a hyperbola (Eq. (8)). In a similar
way, each domain of optimality of a given racquet type is bounded by two parallel lines
µ = µj−1 and µ = µj and one or more pieces of hyperbolae (Eq. (8)). Thus, the boundary
between toroids and racquets is a piecewise curve made of hyperbolic intervals connecting
triple points. The latter are either a toroid triple point where a straight line λ = λi comes
to, or a racquet triple point where two hyperbolae meet a straight line µ = µj. The first
triple point A (µ = µ11, λ = λ1 = 2π√3 + c) is special, it corresponds to the equality of
energies of the single- and double-covered circles and the simplest racquet.
As we have already computed the sequences {Ni}, {Mj} and the corresponding critical
lengths, we may construct the toroid-racquet boundary by applying the following algorithm.
Suppose we know the boundary that separates the NI-toroids and MJ -racquets, that is a
piece of the hyperbola that begins at a point (λI−1, µ⋆), µ⋆ ≥ µJ−1, or (λ⋆, µJ−1), λ⋆ ≥ λI−1,
depending on which coordinates are greater (both points satisfy Eq. (8) for N = NI, M =
MJ ). To fix the other end of our hyperbolic piece we have to choose from two candidates:
(1) (λI, µ⋆⋆) or (2) (λ⋆⋆, µJ) (again both points satisfy Eq. (8) for N = NI, M = MJ ).
If λI < λ⋆⋆ (and µ⋆⋆ < µJ ), then the next triple point is toroidal (1) and the hyperbolic
piece ends there, otherwise the end is the racquet triple point (2). Then, to describe the
next hyperbola piece by Eq. (8), we switch either to NI+1 and MJ in case (1) or to NI and
MJ+1 in case (2). Clearly, we can iterate this procedure to build the boundary curve up to
arbitrary length (Figs. 7, 8).
The values of λi-s depend on the parameter c so that for particular values of the latter
some triple points of different types may coincide. It looks on the diagram of state (Fig. 6)
like four boundary curves converge in one point B where double- and triple-toroids and single-
and double-racquets meet all together. Actually there are a pair of triple points which are
located indistinguishably close to each other because of the value of the parameter c chosen.
The quadruple point only exists for a slightly different value c = 1.4830.
13
B
A
FIG. 6. The diagram of state in the plane (µ, λ). The domains where the conformations shown
are ground states are painted in different colours. The bisector ν ≡ λ/µ = 1 (dashed line) marks
the case of invariant attraction forces independent of the monomer sequence. The upper triangle
ν < 1 corresponds to stronger attraction for the toroidal parallel arrangement of a polymer chain.
The lower triangle ν > 1 contains domains where racquets are ground states because their special
geometry provides enhanced attraction. The relative torsional stiffness is fixed c = 1.5.
As we see in Figs. 7, 8, the boundary between the toroids and racquets remains under the
bisector for given value of the relative torsional stiffness c. We conclude then that if there
is no difference between the interaction energy for parallel and antiparallel alignment, then
the racquet packing always loses to toroids. Figure 9 shows profiles of energies ut (black)
and ur (white) (Eqs. (4) and (5), resp.) for ν = 1 (the bisector at the diagram of state
14
FIG. 7. The diagram of state in the plane (µ, λ) for larger length (c = 1.5). Horizontal lines
separate the toroids and vertical the racquets. Triple points are marked on the toroid-racquet
bondary curve. The dashed line is the bisector ν ≡ λ/µ = 1.
in Figs. 6 -- 8), the regions above the energy curves are painted yellow for toroids and blue
for racquets to clearly demonstrate the difference. Note that though the torsional stiffness
of DNA has not been measured with satisfactory precision6, the variation of c seems to be
within the range of ∼ 0.5 to ∼ 2.5. The boundary between the toroids and racquets remains
below the bisector for such c. However, if there exists another semi-flexible polymer with
stronger torsional stiffness, then the racquets may have lower energy than toroids for some
range of lengths.
Indeed, as seen in Figs. 10, 11, this occurs for c = 20, but only for a
limited interval of lengths: λ / 800. Still, even for such a large c, the energy of racquets is
15
FIG. 8. The diagram of state in the plane (µ, λ) (c = 1.5) (a blowup of Fig. 7).
not significantly lower than for toroids.
IV. DISCUSSION
The diagram of state is in agreement with the earlier conclusion that, in the zero-thickness
model, the racquet-like shapes can be only metastable, i.e. they can only possess slightly
greater energy than the toroidal-like conformations. For open polymer shapes this was shown
in Ref.4. Note that the 1-racquets (or hairpins) can turn out to be stable if the polymer (or
a filament) has a significant thickness that cannot be ignored as for the chromatin fibre17.
However, as we have seen, the situation changes if the polymer interacts with itself differently
16
FIG. 9. Profiles of energies ut (black) and ur (white) (Eqs. (4) and (5), resp.) for c = 1.5, ν = 1.
The regions above the energy curves are painted yellow for the toroids and blue for the racquets.
depending on whether the sequences of the contacting base pairs are homologous or not.
This sequence-dependent effect has been studied for DNA18.
The circular closedness constraint affects both the toroidal and racquet conformations.
Indeed, the racquets become less favourable because the polymer must make even number
of end loops, one loop more than in the minimal open racquet and each loop costs extra
bending energy. On the other hand, the non-zero writhe implies extra torsional energy for
toroids. The latter does not happen if the DNA is nicked and its linking number is not kept
constant. Setting zero torsional stiffness to model this case does not significantly change the
diagram of state (Figs. 6 -- 8).
17
FIG. 10. Profiles of energies ut (black) and ur (white) (Eqs. (4) and (5), resp.) for c = 20.0, ν = 1.
The regions above the energy curves are painted yellow for the toroids and blue for the racquets.
Another problem where a constrained semi-flexible polymer may form toroids is DNA
under external tension in the presence of multivalent ions or polypeptides in the microma-
nipulation experiments14,19 -- 21. It was found that, depending on the enviroment, toroids can
compete with plectonemes.
We can estimate the condensation length for the B-form of DNA. Its bending stiffness
B = kBT lp, where lp ≈ 50 nm. The DNA-DNA interaction energy per length can be
estimated as γ ≈ 0.1kBT nm−16. Thus, the DNA condensation length Lc = pB/γ ≈ 22
nm at room temperature. For c = 1.5, the length that corresponds to the first triple point
A is L = λ1Lc ≈ 300 nm. Thus the shortest DNA for which we expect the stable racquet is
18
FIG. 11. Energies of the toroids ut (yellow) and the racquets ur (blue) (Eqs. (4) and (5), resp.)
for c = 20.0, ν = 1, for longer polymers.
about 103 bp long. The parameter ν equals ∼ 1.7 which is rather high, but smaller values
are sufficient for the racquets made of longer DNA.
If the interacting parts happen to be homologous, then their interaction may be twice
as strong as for nonhomologous molecules18. This effect can be used to make the racquet
shapes stable. To this aim, the sequence may be chosen such that the homologous pieces
come into contact in the particular racquet configuration but not in toroids. Thus we would
have γr = ν2γ = 2γ. The effective racquet condensation length then shortens by factor ν
and the normalised length µ increases by the same factor. This effect is expected to be not
so pronounced for multicovered conformations because the sequence correlations less affect
19
the interaction energy in the tightly packed bundles18.
As we can see on the diagram of state, there is no chance for racquets if ν2 < 1 which
describes the situation when the attraction is enhanced for parallel orientations of the con-
tacting dsDNA intervals. This effect could be easily achieved by building the closed DNA
from homologous pieces joined so that their sequence orientation is same.
On the other hand, if one of the two parts is reversed, then this would give the preference
to the antiparallel alignment like in the handle of the 1-racquet and we have to look into
the part of the diagram for ν2 > 1.
V. CONCLUDING REMARKS
The model considered in this paper assumes that the self-interacting sites tend to arrange
themselves in a parallel bundle. This is an idealisation that cannot be achieved in reality
because of topological constraints22. Nevertheless in many cases DNA forms highly-organised
hexagonal bundles23. This justifies the parallelism approximation.
The proposed tuning of the primary structure of dsDNA may lead to creation of intrinsi-
cally bent pieces, e.g. the so-called A-tracts. This is known to influence the size of toroidal
condensates24 and may also shift the balance between toroids and racquets because the com-
peting conformations differ in curvature. The latter is constant for toroids in the infinitely
thin approximation but varies for racquets.
The dsDNA-dsDNA interaction potential is rather complex and it may have its minima
for a relative orientation of the duplex axes at a non-zero skew angle which depends on
the distance between the molecules6. This may cause formation of twisted bundles even
for unconstrained polymers25. In the simplest case of a single pair of interacting molecules
the complexity of the interaction forces may lead to emergence of a plectonemic structure
which could cause significant change of writhe. This effect can interfere with possible initial
torsional strain of the closed DNA in its reference zero-writhe conformation26. To release
this strain DNA may form plectonemes which could be facilitated or hindered by complex
self-interaction forces. Analysis of these effects is left for future work on elaboration of the
model.
The present model uses a rough estimate of topological features of two series of confor-
mations focusing on the basic difference between two folding patterns. The metastability of
20
the racquet-like shapes and a possibility to finely control their conformational energy may
be exploited to construct a trigger mechanism which will be sensitive to small changes in
the primary structure.
ACKNOWLEDGMENTS
The author thanks G.H.M. van der Heijden, A. Korte, A.A. Kornyshev, D.J. Lee and
R. Cortini for letting him bounce ideas off their bright minds. Support of the UK's Engi-
neering and Physical Sciences Research Council under grant no. EP/H009736/1 is gratefully
acknowledged. The colour palette in Fig. 6 is inspired by Wassily Kandinsky's "Composition
IX" (1936), Mus´ee National d'Art Moderne, Centre Georges Pompidou, Paris.
REFERENCES
1V. A. Bloomfield, Current Opinion in Structural Biology 6, 334 (1996).
2A. G. Cherstvy, Phys. Chem. Chem. Phys. 13, 9942 (2011).
3B. Schnurr, F. C. MacKintosh, and D. R. M. Williams, Europhysics Letters 51, 279
(2000).
4B. Schnurr, F. Gittes, and F. C. MacKintosh, Phys. Rev. E 65, 061904 (2002).
5F. B. Fuller, Proc. Natl. Acad. Sci. USA 68, 815 (1971).
6A.
A. Kornyshev,
D.
J.
Lee,
S.
Leikin,
and
A. Wynveen,
Reviews of Modern Physics 79, 943 (2007).
7J. C. Sitko, E. M. Mateescu, and H. G. Hansma, Biophysical Journal 84, 419 (2003).
8T. Sakaue and K. Yoshikawa, The Journal of Chemical Physics 117, 6323 (2002).
9A. Montesi, M. Pasquali, and F. C. MacKintosh, Phys. Rev. E 69, 021916 (2004).
10V. Barsegov and D. Thirumalai, The Journal of Physical Chemistry B 109, 21979 (2005).
11C. Bottcher, C. Endisch, J.-H. Fuhrhop, C. Catterall, and M. Eaton, Journal of the
American Chemical Society 120, 12 (1998).
12P. Englebienne, P. A. J. Hilbers, E. W. Meijer, T. F. A. De Greef, and A. J. Markvoort,
Soft Matter 8, 7610 (2012).
13E. L. Starostin, The Journal of Chemical Physics 129, 154104 (2008).
21
14C. Battle, B. van den Broek, M. C. Noom, J. van Mameren, G. J. L. Wuite, and F. C.
MacKintosh, Phys. Rev. E 80, 031917 (2009).
15A. E. H. Love, A Treatise on the Mathematical Theory of Elasticity, reprinted Dover, New
York, 1944 ed. (Cambridge University Press, 1927).
16G. H. Wannier, Phys. Rev. 79, 357 (1950).
17B. Mergell, R. Everaers, and H. Schiessel, Phys. Rev. E 70, 011915 (2004).
18A.
G.
Cherstvy,
A.
A.
Kornyshev,
and
S.
Leikin,
The Journal of Physical Chemistry B 108, 6508 (2004).
19B. v. d. Broek, M. C. Noom, J. v. Mameren, C. Battle, F. C. MacKintosh, and G. J. L.
Wuite, Biophysical Journal 98, 1902 (2010).
20J. F. Marko and S. Neukirch, Phys. Rev. E 85, 011908 (2012).
21D. Argudo and P. K. Purohit, Biophysical Journal 103, 118 (2012).
22E. L. Starostin, Journal of Physics: Condensed Matter 18, S187 (2006).
23F. Livolant and A. Leforestier, Progress in Polymer Science 21, 1115 (1996).
24M. R.
Shen,
K. H. Downing,
R. Balhorn,
and N. V. Hud,
Journal of the American Chemical Society 122, 4833 (2000).
25G. M. Grason, Physical Review E 79, 041919 (2009).
26C. Ma and V. A. Bloomfield, Biophysical Journal 67, 1678 (1994).
22
|
1405.7013 | 4 | 1405 | 2015-02-02T16:40:44 | Deconvoluting chain heterogeneities from driven translocation through a nano-pore | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.BM"
] | We study translocation dynamics of a driven compressible semi-flexible chain consisting of alternate blocks of stiff ($S$) and flexible ($F$) segments of size $m$ and $n$ respectively for different chain length $N$ in two dimension (2D). The free parameters in the model are the bending rigidity $\kappa_b$ which controls the three body interaction term, the elastic constant $k_F$ in the FENE (bond) potential between successive monomers, as well as the segmental lengths $m$ and $n$ and the repeat unit $p$ ($N=m_pn_p$) and the solvent viscosity $\gamma$. We demonstrate that due to the change in entropic barrier and the inhomogeneous viscous drag on the chain backbone a variety of scenarios are possible amply manifested in the waiting time distribution of the translocating chain. These information can be deconvoluted to extract the mechanical properties of the chain at various length scales and thus can be used to nanopore based methods to probe bio-molecules, such as DNA, RNA and proteins. | physics.bio-ph | physics |
Deconvoluting chain heterogeneities from driven translocation through a nano-pore
Department of Physics, University of Central Florida, Orlando, Florida 32816-2385, USA
Ramesh Adhikari and Aniket Bhattacharya∗
(Dated: June 11, 2021)
We study translocation dynamics of a driven compressible semi-flexible chain consisting of alter-
nate blocks of stiff (S) and flexible (F ) segments of size m and n respectively for different chain
length N in two dimension (2D). The free parameters in the model are the bending rigidity κb
which controls the three body interaction term, the elastic constant kF in the FENE (bond) poten-
tial between successive monomers, as well as the segmental lengths m and n and the repeat unit p
(N = mpnp) and the solvent viscosity γ. We demonstrate that due to the change in entropic barrier
and the inhomogeneous viscous drag on the chain backbone a variety of scenarios are possible amply
manifested in the waiting time distribution of the translocating chain. These information can be
deconvoluted to extract the mechanical properties of the chain at various length scales and thus can
be used to nanopore based methods to probe bio-molecules, such as DNA, RNA and proteins.
PACS numbers: 87.15.ap, 82.35.Lr, 82.35.Pq
Polymer translocation (PT) through a nano-pore (NP)
is being explored for more than a decade as a NP based
device has the potential to provide single molecule detec-
tion when a DNA is driven electrophoretically through
a NP [1, 2]. Unlike traditional Sanger's method [3] this
does not require amplification; thus one can in principle
analyze a single genome [4]. Progress towards this tar-
get offers challenges to overcome which have attracted a
lot of attention from various disciplines of sciences and
engineering [5, 6]. A large fraction of theoretical and nu-
merical studies have been devoted to translocation stud-
ies of flexible homo-polymers [1, 2]. However, to extract
sequence specific information for a DNA or a protein,
as they translocate and/or unfold through a nanopore,
one needs generalization of the model to account for how
different segments of the translocating polymer interact
with the pore or the solvent. Translocation of the het-
erogeneous polymer has been studied in the past for a
fully flexible polymer where different segments encounter
different forces [7 -- 10]. For periodic blocks one observes
novel periodic fringes from which information about the
block length can in principle be readily extracted [7, 8].
Recently, de Haan and Slater [11] have studied transloca-
tion of rod-coil polymer through a nanopore in the quasi-
static limit (weakly driven through narrow pore and neg-
ligible fluid viscosity). They have used incremental mean
first passage time (IMFPT) [12] approach and verified
that in the quasi-static limit the stiff and flexible seg-
ments can be discriminated due to local entropic mis-
match between the stiff and flexible segments reflected
in the steps and plateaus of the IMFPT of different seg-
ments.
In this letter we provide new insights for the driven
heterogeneous PT through a NP where heterogeneity is
introduced by varying both the bond bending as well as
the bond stretching potentials. We study the transloca-
tion dynamics in the presence of large fluid viscosity and
strong driving force so that the system is not in the quasi-
static limit as in Ref. [11]. Our studies are motivated by
the observation that many bio-polymers, such as DNA
and proteins exhibit helical and random coil segments
whose elastic and bending properties are very different,
so is the entropic contribution due to very different num-
ber and nature of polymeric conformations.
It is also
likely that a double stranded (ds) DNA can be in a par-
tially melted state whose coarse-grained description will
require nonuniform bond bending and bond-stretching
potentials for different regions. As a result, if one wants
to develop a NP based device to detect and identify the
translocating segments, a prior knowledge of their resi-
dence inside the pore will be extremely useful. Naturally,
the length scale of the heterogeneity ξ(n, m), where m
and n are the lengths of the stiff and flexible segments
respectively in each block, will obviously be an important
parameter for the analysis of the translocation problem.
Thus, we first show that a proper coarse graining of the
model in units of ξ will lead to the known results for
the homopolymer translocation. Then we further ana-
lyze the results at the length scale of the blob size ξ
and show how the chain elasticity and the chain stiffness
introduce fine prints in the translocation process. We
explain our findings using Sakaue's non-equilibrium ten-
sion propagation (TP) theory [13] recently verified for
a CG models of semi-flexible chain by us [14 -- 16]. We
have used Lennard-Jones (LJ), Finitely Extensible Non-
linear Elastic (FENE) spring potential and a three body
bond bending potential to mimic excluded volume (EV),
bond stretching between two successive monomers, and
stiffness of the chain respectively, and applied a constant
external force (Fext = 5.0) at the pore in the translo-
cation direction. We have used the Brownian dynamics
(BD) scheme to study the heterogeneous PT problem.
The details of the BD methods are the same as in our
recent publications [14, 15].
Initially we keep the elas-
tic spring constant (kF ) to be the same throughout the
chain and choose the bending stiffness κb = 0 and 16.0
for the fully flexible and the stiff segments respectively.
Later we show that by making the elastic potential for
the relatively more flexible part weaker one can reverse
the relative friction on the chain segments which results
in novel waiting time distributions serving as the finger-
print of the structural motifs translocating through the
pore.
2
FIG. 1. Blob model of a polymer chain of chain length (N =
24) and segmental length (m = 4). Each repeated unit can
be considered as a single blob of length ξ ∼ mβ. Please see
the text below.
• Blobsize and scaling: We consider heterogeneous
chains consisting of alternate symmetric (m = n) peri-
odic blocks of stiff and flexible segments of m monomers
so that the block-length is 2m (m = 1,2,3,4) as shown
in Fig. 1. First we investigate how the alternate stiff
and flexible segments of equal
length affect the end-
to-end distance hRN (m)i and the mean first passage
time (MFPT) as a function of the periodic block-length
(Fig. 1), compared to a homo-polymer of equal contour
length N . To a first approximation one can think of this
chain as a flexible chain of N/2m segments, of certain
blob size ξ. The blob size ξ in general will be a function
of the block-length and bending rigidity of the flexible
and stiff segments. For our particular choice of bending
rigidity for the flexible (κb = 0) and stiff (κb = 16) seg-
ments from simulation results for N = 64 − 256 we find
an expected power law scaling ξ ∼ mβ where β = 0.87
(Fig. 2). Obviously the exponent β is non-universal as
it depends on κb and kF , but the universal aspects of
the entire chain can be regained through scaling with ξ
as shown in Fig. 2. The conformation statistics of this
basic unit ξ controls both the conformation and translo-
cation properties of the entire chain as follows. We can
write hRN i ≡ hpR2
N i ∼ hξi(N/2m)ν ∼ mβN ν /mν,
where ν is the Flory exponent. This implies hRN i/N ν ∼
mβ−ν = m0.12 (where ν = 0.75 is the Flory exponent
in 2D). Simulation data in the insets of Fig. 2(a) con-
firms our scaling prediction. Likewise, we show that the
MFPT hτ i/N 2ν ∼ m0.09. For small N it has been found
earlier that hτ i ∼ hRN i/N −ν ∼ N 2ν [17]. Therefore,
as expected by proper coarse graining by the elemental
block we get back the results for the fully flexible chain.
We now show how characteristics of translocation are af-
fected by the chain heterogeneity.
• Effect of chain heterogeneity on translocation: In pre-
FIG. 2. (a) Log-log plot of blob-size hξi as a function of m
for N = 64 (black plus), N = 128 (violet cross) and N = 256
(orange right-triangle). The solid line represents hξi ∼ m0.87.
Insets: (i) log-log plot of hRN i as a function of N for different
m, (ii) collapse of hRN i/m0.12 ∼ N ν on the same master
plot. (b) Log-log plot of hτ i as a function of N for different
m. Inset: scaling and collapse of hτ i/m0.09 ∼ N 2ν .
senting the results we use the notation (FmSn)p/(SmFn)p
to denote p blocks of an ordered flexible/stiff and
stiff/flexible segments of length m and n respectively
(N = (m + n)p) and that the flexible/stiff segment enters
the pore first. Fig. 3 and Fig. 4 reveal quite a few novel
results that we explain using TP theory. For small block-
length the order in which the chain enters the pore (either
stiff or flexible segment) neither make a big difference in
the shape of the histogram (Fig. 3(a)) nor in the MFPT
(Fig. 4). For larger block lengths the difference between
the histograms for SmFm and FmSm are quite clear and
the dependence of τ on m are also different as seen in
Fig. 4. For the case when the stiff portion enters the
pore first, the MFPT monotonically increases but in the
other case it shows a maximum (Fig. 4). We now explain
this in terms of our recent analysis of the translocation
of semi-flexible chain using TP theory where we showed
that a stiffer chain takes a longer time to translocate [14 --
16]. When the block lengths are small, TP gets intermit-
tently hindered as the tension propagates through alter-
nate stiff and flexible regions. For longer blocks tension
can propagate more effectively unhindered for a longer
time. Therefore, when a long stiff segment enters the
pore first it increases the MFPT. But, when a long flexi-
ble segment enters the pore first it decreases the MFPT.
This results a maximum in the hτ i/hτ i0 vs. m curve for
the FS orientation. The difference of MFPT for SmFm
and FmSm becomes maximum when m = N/2. For rel-
atively longer block lengths it makes a big difference in
MFPT.
3
FIG. 3. Histograms of first passage time for chain length
N = 128 and segmental length (a) m = 4, (b) m = 16, (c) m
= 32, and (d) m = 64. The dotted/solid lines represent the
flexible(FmSm)/stiff(SmFm) segment entering the pore first.
For larger block size the effect of order of entry is clearly
visible.
FIG. 4. MFPT (scaled by the MFPT of respective flexible
homo-polymer) for chains FmSm and SmFm as a function of
m/N for chains N = 64 (green diamonds) and N = 128 (blue
up-triangles). The open/closed symbols correspond to flexi-
ble/stiff segment entering the pore first. The inset shows the
ratio of the MFPT for SF to F S orientation. The nanopore
is capable of differentiating if a flexible (F) block or a stiff (S)
block entered the pore first.
• Waiting time distribution: The total time spent by
a monomer inside the pore is defined as its waiting time
W (s), where s is the index of monomer inside the pore
(translocation coordinate). Sum of the waiting time for
all monomers is the MFPT i.e. PN
s=1 W (s) = hτ i. The
effect of TP in stiff and flexible parts becomes most vis-
ible in the waiting time distribution of the individual
monomers of the chain as shown in Fig. 5. We notice that
the envelopes for the corresponding homo-polymers for a
fully flexible chain (κb = 0, solid orange line) and for the
stiffer chain (κb = 16, solid green line) respectively serve
as bounds for the heterogeneous chains [22]. As explained
in our previous publication [15] the TP time corresponds
to the maximum of these curves and shifts toward a lower
s value for a stiffer chain. Bearing this in mind we can
FIG. 5. Waiting time distribution for a N = 128 chain with
the block-length (a) 16 and (b) 32. Azure (open circles) and
Blue (filled squares) correspond to the flexible and stiff seg-
ments when the flexible segment enters the pore first (FmSm).
Magenta (open circles) and Red (filled squares) correspond to
the flexible and stiff segments when the stiff segment enters
the pore first (SmFm). The solid green and orange lines corre-
spond to the waiting time distributions for the corresponding
stiff (κb = 16.0) and fully flexible (κb = 0.0) homo-polymers
respectively.
reconcile the fringe pattern in the light of the TP theory.
The pattern has the following features: (i) The number
of fringes is equal to the number of blocks. This is be-
cause on an average stiffer portions take longer time to
translocate. (ii) The fringes for SmFm and FmSm are out
of phase for the same reason. (iii) The chain heterogene-
ity affects the waiting distribution most at an early time;
beyond the largest TP time (i.e., the peak position of the
envelope for κb = 0) the waiting time of the individual
monomers (excepting which are at the border separating
the stiff and flexible segments) becomes identical to that
of the corresponding homogeneous chain. This again ex-
emplifies to analyze the driven translocation as a pre and
post TP events. Please note that the maxima of the W (s)
for the heterogeneous chain lie in between the maxima for
the corresponding homogeneous cases.
• Effect of friction and driving force: In Fig. 5 we chose
a value of the solvent friction associated to each monomer
γ = 0.7 for which we find that a stiffer segment translo-
cates slower through the pore. We now discuss how a
variation of the solvent friction will affect this conclu-
sion. We first show that the MFPT of a homopolymer of
certain length exhibits a crossover as one varies the sol-
vent viscosity (Fig. 6(a)). It is only for extremely small γ
(quasi-static limit) the stiffer segment translocates faster
as studied in [11]. We also have reproduced the result for
a particular set of parameters (black line in Fig. 6(b)).
We have shown 3D (instead of 2D) data in Fig. 6(b) and
(c) only for better resolution. This crossover effect can
be explained using Sakaue's tension propagation (TP)
4
monomers.
In quasi-static limit, significantly larger local entropic
barrier of a "coil" segment causes longer residence time.
This effect is reflected as steps in the IMFPT (Fig. 6(b))
and peaks in the waiting time distribution (Fig. 6(c)).
But for the non-equilibrium situation, when the stiffer
segment enters the pore, tension propagates faster along
the chain backbone [13, 15] and more monomers in the
cis-side set in motion. For large solvent friction this may
produce larger viscous drag dominating over local en-
tropic barrier resulting in the stiffer segments translo-
cating slower than the flexible segment.
In this case
the peaks in the waiting time distribution disappear
(red color in Fig. 6(c)). Accordingly, one sees qualita-
tive changes in the corresponding IMFPT (red color in
Fig. 6(b)). Therefore, the relative fast/slow translocation
of rod /coil segments through the nanopore depends on
the relative values of pore friction, solvent friction, and
applied bias.
• Heterogeneous chain with a variable spring con-
stant: Finally we have extended these studies to see the
consequences of allowing the elastic potential between
the successive beads to be different in each block. This
situation may occur when individual building blocks are
connected by linkers of different elasticity. Fig. 7 shows
the various combination of the spring constants kF for the
heterogeneous chain. The first four graphs Fig. 7(a)-(d)
correspond to the waiting time distribution for the chain
with equal number of monomers in each of the flexible
and stiff segments. Fig. 7a is the graph where all the F
and S segments have the same kF = 100 qualitatively
similar to Fig. 5. In Figs. 7(a)-(d) one can see the effect
of reduced value of kF for the flexible portion only.
Figs. 7(e)-(f) represent the waiting time distributions
for the unequal length of the flexible and stiff segments.
The flexible segment, being shorter, looses the conforma-
tional entropic height but the contribution of the FENE
force in the direction of translocation is enhanced. We
can see the effect of this enhancement in the increased
back and forth motion (low frequency phonons of larger
amplitude to softer bonds) of the chain towards the
translocation direction. The smaller is the value of kF
the larger will be the amplitude of the phonons mode
which results in a longer translocation time. Therefore,
when we reduce the strength of the FENE interaction for
the coil, the coil translocates slower and we got the wait-
ing time distribution picture inverted for the stiff and
flexible segments as seen from a comparison of Fig. 7(a)
to Fig. 7(d). This will be most prominent if the stiff
segments were chosen as rigid rods.
Fig. 7(c)-(f) show the end monomer of each semi-
flexible segment has a larger waiting time. This indicates
a larger barrier height for the flexible segments. Once the
barrier is overcome by the first monomer of the flexible
segment, all the following monomers of the flexible seg-
ments pass through the pore faster. The end monomer
(a) The MFPT for flexible and semi-flexible ho-
FIG. 6.
mopolymers of length N = 64 as a function of solvent-
monomer friction γ.
(b) The IMFPT and (c) the waiting
time distribution as a function of s-coordinate for a chain
(N = 70) in 3D with four stiff segments (κb = 100) each of
length (m = 10) and five flexible segments (κb = 0) each of
length (n = 6) provided that (for (b) and (c)) the first flexible
segment is already in the trans-side at t = 0.
theory [13]. When we use a larger value of γ (implying
stiffer segment translocates slower) and/or a bias F the
IMFPT changes qualitatively (Fig. 6(b)), which is more
prominently seen in the waiting time distribution of the
individual monomers (Fig 6(c)).
Using formulae for solvent friction from the bulk
Γsolv = γN ν and pore friction Γpore ∼ Apore
d−1 + pγ which
have been discussed in Ref. [18, 19] we have checked that
γ = 0.7 and γ = 0.1 (for the chain lengths used in our
simulation) correspond to solvent dominated and pore
friction dominated regimes respectively. At high Γsolv,
de-Haan and Slater showed that the MFPT increases lin-
early with γ [20] for a fully flexible chain. We see the
same trend to be valid also for semi-flexible chains, al-
beit beyond a critical value (Fig. 6(a)). But at low Γsolv,
the dependence of MFPT on γ becomes non-monotonic
and it exhibits a minimum for γ = γm [21]. This γm
marks the onset of change in the qualitative behavior of
IMFPT or the waiting time distribution of the individual
10
5
10
5
10
5
10
5
10
5
10
5
0
0
)
s
(
w
m:n=1:1
(kF)coil= 100, (kF)rod = 100
m:n=1:1
(kF)coil= 40, (kF)rod = 100
m:n=1:1
(kF)coil= 10, (kF)rod = 100
m:n=1:1
(kF)coil= 5, (kF)rod = 100
m:n=5:3
(kF)coil= 5, (kF)rod = 100
m:n=3:1
(a)
(b)
(c)
(d)
(e)
(f)
(kF)coil= 5, (kF)rod = 100
32
64
s
96 128
5
compensated by the waiting time for a flexible segment
but having softer elastic bonds. This observation can
be exploited to tune to control the passage of polymers
through NP. It is interesting to note from Fig. 7 that the
variation in waiting time distribution arising out of the
bending stiffness variation and bond length variation can
be differentiated. Therefore, these patterns can serve as
references to characterize structural heterogeneity of an
unknown polymer translocating through a nanopore. We
hope the results reported in this letter will be helpful in
deciphering translocating characteristic of bio-polymers
observed experimentally.
This research has been partially supported by a UCF
College of Science Seed grant. We thank Profs. Gary
Slater and Hendrick de Hann for useful discussions.
FIG. 7. The waiting time distribution as a function of s-
coordinate for a chain (N = 128) with variable kF and stiff-
flexible segmental length ratio (m/n). The bending stiffness
κb for flexible (red circles) and stiff (blue squares) segments
are 0 and 16 respectively. The elastic stiffness (kF ) is 100 for
stiff segments [(a)-(f)]. For flexible segments (a) kF = 100 (b)
kF = 40 (c) kF = 10 (d) kF = 5 (e) kF = 5 and (f) kF = 5.
The stiff and flexible segments are of equal length except in
(e) m : n = 5 : 3 and (f) m : n = 3 : 1.
of the flexible segment and the first monomer of the stiff
segment have the lowest waiting time which means that
they have negligible barrier to overcome. Furthermore a
visual comparison of Fig. 5 and Fig. 7 shows that the ori-
gin of the details of the waiting time distributions possi-
bly be differentiated by a spectral decomposition analysis
of the waiting time distribution.
To summarize, we have demonstrated how a nanopore
can sense structural heterogeneity of a bio-polymer
driven through a nanopore. Not only do monomers be-
longing to the flexible and stiff part exhibit different wait-
ing time distributions, we have also demonstrated how a
nanopore can sense which end of the polymer enters the
pore first. Translating this information for a dsDNA will
imply that the nanopore can differentiate the 3-5 or 5-3
ends of a translocating DNA. We have explained these
results using the concepts of TP theory. We have clearly
demonstrated how the fluid viscosity and an external
bias can affect the relative speed of the stiff and flexible
segments. Furthermore, unlike previously reported stud-
ies [11] we, for the first time, analyzed the interplay of the
effects of polymer heterogeneity caused by the variation
of elastic and bending stiffness. We have demonstrated
that softer elastic bonds raise the MFPT [23]. Therefore,
an increase in waiting time for a stiff segment can be
∗ Author to whom the correspondence should be ad-
dressed; [email protected]
[1] M. Muthukumar Polymer Translocation (CRC Press,
Boca Raton, 2011).
[2] A. Milchev, J. Phys. Condens. Matter 23, 103101 (2011).
[3] F. Sanger and A. R. Coulson, J. Mol. Biol. 95, 441 (1975);
F. Sanger, S. Nicklen, and A. R. Coulson, Proc. Natl.
Acad. Sci. USA 74, 5463 (1977).
[4] E. R. Mardis, Nature 470, 198 (2011).
[5] L. Movileanu, Soft Matter 4, 925 (2008).
[6] D. Rodriguez-Larrea and H. Bayley, Nat. Nanotech. 8,
288 (2013).
[7] K. Luo, T. Ala-Nissila, S.C. Ying and A. Bhattacharya,
J. Chem. Phys. 126, 145101 (2007).
[8] K. Luo, T. Ala-Nissila, S.C. Ying and A. Bhattacharya,
Phys. Rev. Lett. 100, 058101 (2008).
[9] M. G. Gauthier and G. W. Slater, J. Chem. Phys. 128,
175103 (2008).
[10] S. Mirigan, Y. Wang, and M. Muthukumar, J. Chem.
Phys. 137, 064904 (2012).
[11] H. W. de Hann and G. W. Slater, Phys. Rev. Lett. 110,
048101 (2013).
[12] H. W. de Hann, G. W. Slater, J. Chem. Phys. 134,
154905 (2011).
[13] T. Sakaue, Phys. Rev. E 76, 021803 (2007); ibid 81,
041808 (2010).
[14] A. Bhattacharya, J. Polymer Science Series C 55, 60-69
(2013).
[15] R. Adhikari and A. Bhattacharya, J. Chem. Phys. 138,
204909 (2013).
[16] Sakaue's TP picture was demonstrated to be valid for a
fully flexible chain by translating the original theory to a
BD scheme [18, 24]. In refs. [14, 15] we discussed how the
TP picture will be affected by the stiffness of the chain.
[17] K. Luo, I. Huopaniemi, T. Ala-Nissila, and S.-C. Ying,
J. Chem. Phys. 124, 114704 (2006).
[18] T. Ikonen, A. Bhattacharya, T. Ala-Nissila and W. Sung,
J. Chem. Phys. 137, 085101 (2012).
[19] T. Ikonen, A. Bhattacharya, T. Ala-Nissila, and W.
Sung, Europhys. Lett. 103, 38001 (2013).
[20] H. W. de Haan and G. W. Slater, J. Chem. Phys. 136,
154903 (2012)
[21] R. Adhikari and A. Bhattacharya, Unpublished.
In the pore friction dominated regime upon decreasing
the solvent viscosity the system is closer to its equilibrium
state and the translocation dynamics is determined by
entropic barrier crossing. This entropic barrier is smaller
for a stiffer chain and becomes negligible for rod-like poly-
mer. At low viscosity, the force due to pore friction in-
creases as the solvent viscosity decreases resulting in an
increase in the MFPT which explains the minimum of
the MFPT at a particular viscosity γ ∼ γm.
6
[22] This characteristic pattern of waiting time distribution
seems to be a generic feature as it was seen earlier for
segments under different bias [7] or due to different pore-
segment interaction [8].
[23] We have simulated driven translocation of a one dimen-
sional compressible chain and find that the MFPT with
the elastic constant kF increases as hτ i ∼ N 2k−1/8
.
[24] T. Ikonen, A. Bhattacharya, T. Ala-Nissila and W. Sung,
F
Phys. Rev. E 85, 051803 (2012).
|
1001.2524 | 1 | 1001 | 2010-01-14T18:21:46 | Lipid Ion Channels | [
"physics.bio-ph"
] | The interpretation electrical phenomena in biomembranes is usually based on the assumption that the experimentally found discrete ion conduction events are due to a particular class of proteins called ion channels while the lipid membrane is considered being an inert electrical insulator. The particular protein structure is thought to be related to ion specificity, specific recognition of drugs by receptors and to macroscopic phenomena as nerve pulse propagation. However, lipid membranes in their chain melting regime are known to be highly permeable to ions, water and small molecules, and are therefore not always inert. In voltage-clamp experiments one finds quantized conduction events through protein-free membranes in their melting regime similar to or even undistinguishable from those attributed to proteins. This constitutes a conceptual problem for the interpretation of electrophysiological data obtained from biological membrane preparations. Here, we review the experimental evidence for lipid ion channels, their properties and the physical chemistry underlying their creation. We introduce into the thermodynamic theory of membrane fluctuations from which the lipid channels originate. Furthermore, we demonstrate how the appearance of lipid channels can be influenced by the alteration of the thermodynamic variables (temperature, pressure, tension, chemical potentials) in a coherent description that is free of parameters. This description leads to pores that display dwell times closely coupled to the fluctuation lifetime via the fluctuation-dissipation theorem. Drugs as anesthetics and neurotransmitters are shown to influence the channel likelihood and their lifetimes in a predictable manner. We also discuss the role of proteins in influencing the likelihood of lipid channel formation. | physics.bio-ph | physics | Review: Lipid Ion Channels
Thomas Heimburg
Membrane Biophysics Group, Niels Bohr Institute, University of Copenhagen, Denmark
0
1
0
2
n
a
J
4
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
4
2
5
2
.
1
0
0
1
:
v
i
X
r
a
ABSTRACT
The interpretation electrical phenomena in biomembranes is usually based on the assumption that the
experimentally found discrete ion conduction events are due to a particular class of proteins called ion channels while the
lipid membrane is considered being an inert electrical insulator. The particular protein structure is thought to be related
to ion specificity, specific recognition of drugs by receptors and to macroscopic phenomena as nerve pulse propagation.
However, lipid membranes in their chain melting regime are known to be highly permeable to ions, water and small
molecules, and are therefore not always inert. In voltage-clamp experiments one finds quantized conduction events
through protein-free membranes in their melting regime similar to or even undistinguishable from those attributed
to proteins. This constitutes a conceptual problem for the interpretation of electrophysiological data obtained from
biological membrane preparations.
Here, we review the experimental evidence for lipid ion channels, their properties and the physical chemistry underlying
their creation. We introduce into the thermodynamic theory of membrane fluctuations from which the lipid channels
originate. Furthermore, we demonstrate how the appearance of lipid channels can be influenced by the alteration of
the thermodynamic variables (temperature, pressure, tension, chemical potentials) in a coherent description that is free
of parameters. This description leads to pores that display dwell times closely coupled to the fluctuation lifetime via
the fluctuation-dissipation theorem. Drugs as anesthetics and neurotransmitters are shown to influence the channel
likelihood and their lifetimes in a predictable manner. We also discuss the role of proteins in influencing the likelihood of
lipid channel formation.
keywords: lipid pores, phase transitions, fluctuations, anesthetics, recep-
tors, channel proteins, fluctuation-dissipation, electroporation
1
Introduction
Biological membranes mainly consist of a lipid bilayer into
which proteins are embedded. The mass (or volume) ratio
between proteins and lipids ranges between 0.25 (lung sur-
factant) and 4 (purple membrane of halobacteria). Typical,
biomembranes display a protein-lipid ratio of approximately
one. This includes the extra-membraneous parts of the proteins
and surface-associated protein. Therefore, the intra-membrane
parts of the proteins represent a smaller fraction of the central
part of the membrane. Thus, even in densely crowded biologi-
cal membranes most of the in-plane membrane area consists of
lipids.
Textbook models about the electrical transport properties of
membranes consider the lipid bilayer as an electrical insula-
tor. For instance, the Hodgkin- Huxley model of the nervous
impulse considers the membrane as an inert capacitor and
attributes currents to conductance changes occurring in ion
selective protein channels [1]. The activity of protein ion chan-
nels is associated to quantized current events of order of 1-50
pA (pico Amp`eres) at a voltage of order 100mV corresponding
to channel conductances of about 10-500 pS (pico Siemens).
The typical open-dwell time of such channels is of the order of
a few milliseconds.
Interestingly, however, synthetic lipid membranes close to
phase transitions are not inert but are very permeable for small
molecules [2], ions [3, 4, 5, 2] and water [6]. Further, they
display stepwise conductance events very similar to those of
[7, 8, 9, 10, 11, 12]).
proteins, i.e., with similar conductances and life times (some
early publications:
In their appear-
ance, they are practically indistinguishable from recordings of
protein-containing membranes. Although this is surprisingly
little known, it constitutes a severe problem for the investi-
gation of biological membranes and protein channels. These
events will be called lipid channels throughout this manuscript.
The finding of an enhanced lipid membrane permeability close
to transitions is particularly important since biomembranes
seem to exist in a state slightly (about 10-15◦) above melting
transitions. The membrane compositions of fish and bacteria
adapt upon changes in body temperature, pressure or solvent
conditions such that this membrane state is maintained (re-
viewed in [13]).
It seems therefore plausible to assuming a
functional purpose for preserving a particular physical state of
the membrane.
The most common ways to measure conductances of mem-
branes are the black lipid membrane (BLM) technique by
Montal & Mueller [14] and the patch-clamp technique pio-
neered by Neher & Sackmann [15].
In the Montal-Mueller
technique an artificial membrane is formed over a small teflon
hole (normally with a diameter of 50-250 µm). The patch-
clamp technique relies on the measurement of currents through
small membrane patches defined by a glass capillary of a
diameters ranging from 1-10 µm. While the Montal-Mueller
technique is suited for measuring artificial systems of defined
composition, e.g., one single protein species reconstituted into
a single lipid bilayer, the patch-clamp method is mainly used
for recording currents through biological membranes with
complex and often unknown composition. Patch pipettes have
also been used for measurements on synthetic BLMs [16, 17].
1
In both techniques, the observed membrane area is much larger
than that of a protein. The diameter of an ion channel protein,
e.g. the potassium channel, is about 5nm. This implies that the
smallest patch in a patch-clamp experiment is already 40000
times larger than the cross-sectional area of the protein. The
hole in an BLM experiment is rather 400 million time larger
than a typical protein channel area (assuming an aperture of
100 µm). However, the electrophysiological experiment in
itself cannot tell where along which path the ion currents flow.
Thus, if one wants to attribute a particular current event to a
protein one has to assume that at least 99.998% of the sur-
rounding membrane (including the other membrane proteins)
are inert in the electrophysiological experiment. Therefore, it
can generally not be excluded on the basis of the experiment
itself that during an electrophysiological experiments one also
finds currents through the lipid membrane.
In the Hodgkin-
Huxley paper [1], the possibility of leak currents is included.
If these are structureless small amplitude currents in the back-
ground this does not represent a major problem. However, leak
currents represent a significant conceptual problem to protein
channel recordings if they display a similar signature as the
protein events.
The purpose of this review is to demonstrate that in fact
the ion currents through lipid membranes resemble those
through proteins to a degree that they become indistinguish-
able. Their occurrence depends on temperature, lateral tension,
membrane-associated drugs as anesthetics and neurotransmit-
ters, pH, calcium concentration, voltage and other thermody-
namic variables. Here, we review the literature on membrane
permeability and lipid ion channels, and introduce into the
thermodynamics of their creation. In this context we discuss a
possible role for proteins as membrane perturbations that alter
the state of the lipid membrane.
2 Macroscopic changes of permeability
in transitions
2.1 Experiments
During the 1960s and 70s, many researchers were investigating
the permeability of biomembranes for ions. Proteins were
assumed being the major players. However, accurate controls
of the permeability of the lipid membrane for ions was needed
as a control. Hauser et al.
[18] studied the permeability of
lipid extracts from egg yolk or ox brain and found the perme-
ability for ions being small compared to the permeability of
biological membranes. The authors concluded that the role of
the lipid bilayer for permeability is negligible for the case of a
biological membrane suggesting that the predominant part of
permeability has to be attributed to proteins.
However, it was soon shown that the permeability can be
orders of magnitude larger close to the chain melting regime
of lipid membranes. Papahadjopoulos et al. [3] were the first
Figure 1: Left: Permeability of DPPG vesicles for ra-
dioactive sodium, 22N a+, close to the melting transition,
adapted from [3]. Right: Permeability of DMPC mem-
branes for Co2+. Courtesy O.G. Mouritsen, cf. [5]. The
dotted profiles are a gide to the eye. The dotted vertical
line marks the melting transition temperature, Tm.
to demonstrate that the permeability for sodium ions (they
used radiolabeled 22Na+ ions) increased by at least a factor of
100 in the phase transitions of dipalmitoyl phosphatidylglyc-
erol (DPPG) and dipalmitoyl phosphatidylcholine (DMPC)
in agreement with the phase transitions of these lipids as
measured by fluorescence changes of embedded markers. The
permeability curve for DPPG membranes is shown in Fig. 1
(left). The permeation time scales in this paper seem extremely
small (of the order of hours) which might be related to the fact
that none of the data points was recorded at the phase transition
directly. The permeation profile for DPPC was found to be
similar similar (not shown). It was demonstrated that choles-
terol both abolishes the permeability maximum and the chain
melting discontinuity. Along the same lines Sabra et al. [5]
found that the permeability of dimyristoyl phophatidylcholine
(DMPC) membranes for Co2+ was drastically enhanced in
the phase transition regime (Fig. 1, right). These authors also
demonstrated that the insecticide lindane changes the perme-
ability. This phenomenon will be discussed in more detail in
section 6.3.
Jansen and collaborators [6] showed that membranes in their
transition are much more permeable to water (Fig. 2, left).
Vesicles filled with D2O display a contrast with respect to an
H2O background leading to enhanced light scattering in an op-
tical experiment. Permeation for D2O was monitored in a rapid
mixing stopped-flow experiment. It leads to a mixing of H2O
and D2O and a loss of scattering contrast. The authors found
that the permeability of a DMPC membrane changes strongly
in the phase transition regime such that exchange of water from
a vesicle is complete after 2ms, the fastest time that could be
recorded in this experiment. The same finding was reported
for other lipids as DPPC and distearoyl phosphatidylcholine
(DSPC), a series of phosphatidyl ethanolamines, phosphatidyl
serines and phosphatidylglycerols, rendering this study one of
2
the most complete works on membrane permeability. Since
water permeability typically is discussed in the context of aqua-
porines [19] it is important to point out the the membrane itself
is highly permeable to water once one is close to a transition,
which seems to be the case for many biological membranes
[20, 13]. Blicher and coworkers [2] found in fluorescence
Figure 2: Left: Permeation time scale for water. Light
scattering of DMPC vesicles filled with D2O in a H2O
environment is monitored after a fixed time in a stopped-
flow experiment. The decrease of the scattering ampli-
tude close to Tm=23.7◦C is due to fast mixing of D2O
and H2O. Adapted from [6]. The dotted line is a guide to
the eye. Right: The permeation rate of Rhodamine 6G out
of vesicles from a mixture of DPPC and DPPG (95:5) in
the presence of 200mM NaCl. The solid line is the exper-
imental heat capacity profile for vesicles of the identical
preparation. Adapted from [2].
correlation spectroscopy measurements that vesicles mixtures
of DPPC and DPPG became significantly more permeable
for the positively charged fluorescence marker rhodamine 6G
(R6G) in the phase transition regime (Fig. 2, right). While
the permeation time scale in the solid lipid phase was of the
order of many hours, it was 100 seconds and faster in the
transition regime (the fastest time scale that could be recorded
in the FCS experiment). The characteristic dimension of this
marker is approximately 1nm. A control experiment with a
fluorescent-labeled sugar that are larger than R6G did not show
any measurable permeation on the timescale of hours. This
demonstrates that
1. vesicles stay intact and do not rupture in the transition.
One may confidently assume that this is also the case for
the experiments shown in Fig. 1 and 2 (left). 2.
2. sizes of pores responsible for the permeation process
must be of nm-size.
The conduction for ions and small molecules described in this
section was named 'macroscopic' permeability because the ex-
periments were made on ensembles of vesicles. This kind of
measurement describes the relation of the lipid melting events
and the permeability correctly but does not make any statement
3
on the molecular nature of any permeation process. The latter
will be discussed in more detail below in section 4.
3 Techniques for measurement of lipid
channel activity
The electrical properties of Lipid bilayers and biological mem-
branes are measured through small membrane patches defined
by a hole in a hydrophobic substrate [14] (black lipid mem-
branes), the tip of a glass pipette [21] (patch-clamp), or more
recently by a hole in a glass surface on a chip [22] in a technique
called 'planar patch-clamp'. Both Montal-Mueller and patch-
clamp technique have been used to investigate permeability and
channel formation in pure lipid membranes and are briefly de-
scribed below.
3.1 Black lipid membranes / Montal-Mueller
technique
In 1972, Montal and Mueller [14] described a method to form
lipid bilayers formed over a hole in a thin hydrophobic film
(e.g., teflon) from monolayers. Monolayers are formed on an
electrolyte surface from a solvent solution, typically containing
decane, octane, hexane, pentane, chloroform or other apolar
solvents. The monolayer trough is separated be a teflon film
into two compartments. A tiny hole in the film (several 10
µm) created by a needle or by an electric spark slightly is
located slightly above the monolayer surface. The level of the
aqueous buffer is now raised and the monolayers fold over
the aperture in the teflon film as shown schematically in Fig.
3 (left). The monolayers merge into a bilayer at the edge of
the teflon aperture. At these edges one expects considerable
mechanical tension of quite unclear physics. In order to reduce
this tension the teflon aperture is normally pre-painted with
hexadecane, a long-chain alkane that is practically water insol-
uble. Hexadecane must be considered as a impurification that
accumulates at the edge of the aperture. It is quite difficult to
estimate how much hexadecane is dissolved in the membrane.
For this reason, the composition of the membrane is always
somewhat unclear even when the composition of the monofilm
is exactly known. Currents are measured with electrodes in the
two compartments in a voltage clamp experiment. The capac-
itance can be measured by continuously changinging voltage
such that dV /dt = const. The capacitance Cm is obtained
from the capacitive current Ic = Cm dV /dt necessary to
charge the capacitor (assuming that the capacitance itself is a
constant, dCm/dt = 0, which may not necessarily be correct if
voltage changes the state of the membrane). The measurement
of the capacitance is required to ensure that only one single
membrane has formed. Typical capacitances of membranes are
between 0.6 and 1 µF/cm2.
Membranes forming over a hole appear black in visible light
due to destructive interference of light reflected from the
solvents. The interface of the patch pipette and the suction is a
source for perturbation [25].
Patch pipettes can also be used to study BLMs [16, 17]. The
tip of a glass pipette filled with electrolyte solution is placed on
a buffer surface. Lipids dissolved in an organic volatile sol-
vent (we use mixtures of hexane and ethanol) are brought into
contact with the outer surface of the patch pipette. When the so-
lution runs down the pipette a membrane spontaneously forms
at the pipette tip as schematically shown in Fig. 3 (right). Af-
ter membrane formation as evident in the capacitance measure-
ment slight suction is applied. In this technique no long chain
alkanes are used and one can assume that after some equilibra-
tion time the membrane at the pipette tip is practically solvent
free (W. Hanke, personal communication). Hanke and cowork-
ers [17] also describe a variation of this technique called tip-
dipping where the membrane patch is generated from dipping
the pipette tip into monolayer surfaces.
4 Lipid channels
The evidence for pores (channels) within membrane proteins
selectively passing ions rests partially on the observation that
the currents through membranes are quantized (Fig.4a). The
fact that the currents are small and correspond to aqueous
channels of the dimensions of a few A suggests that the con-
duction of ions occurs along objects of molecular dimension.
They have been attributed to particular proteins called ion
channels. The literature about these proteins is huge (e.g.,
[26]). It seems therefore surprising that one can find similar
quantized current events in pure lipid membranes that do not
contain proteins. Fig. 4a shows the historical experiment of
Neher & Sackmann [15] on frog muscle cells containing the
acetylcholine receptor in the presence of an agonist molecule.
Fig. 4b shows the currents through a lipid membrane from
a mixture of dioleoyl phosphatidylcholine (DOPC) and di-
palmitoyl phosphatidylcholine (DPPC) having a chain melting
transition close to the experimental temperature. The melting
temperature is important since it has been shown in section 2
that membranes are most permeable in their melting transition.
It is obvious that the two traces display very similar features
both regarding amplitude and time scales of the currents. In
the following we will characterize the lipid channels in more
detail.
4.1 Early experiments
To the best of our knowledge the first channel events recorded
in pure lipid membranes are from Yafuso et al. in 1974 [7], two
years before the famous paper by Neher and Sakmann on the
acetylcholine receptor shown in Fig. 4a. Their experiments on
bilayers from oxidized cholesterol are shown in Fig. 5 (left).
They found spontaneous conductance changes and multilevel
Figure 3: Two methods to study currents through artifi-
cial membranes. Left: In the Montal-Mueller technique
[14] two monolayers merge into a bilayer over an aper-
ture in a teflon film . Typical aperture size is 50-250 µm.
Right: One can also generate a membrane patch on a
glass pipette tip dipped into the surface of a Langmuir
monolayer trough [16, 17]. Typical pipette tip sizes are
1-10µm. The drawings were adapted from [23]. The two
microscopic images on the bottom show the holes in a
teflon film [24] and a pipette tip used for experiments in
our lab (courtesy Katrine R. Laub, NBI).
two bilayer surfaces. They are called black lipid membranes
(BLMs) because their formation can be seen by visual in-
spection under a microscope. The typical hole diameter in
a Montal-Mueller experiment is 50 to 250 µm (Fig. 3, left
bottom). The disadvantage of this technique is the presence of
solvent impurities and the large tension at the edge of the teflon
aperture. Furthermore, due to the large diameter of the hole the
membranes are often quite unstable. However, this technique
is often used to study lipid membrane permeability but also the
conductance through channel proteins and peptide pores.
3.2 Patch-clamp technique
The patch-clamp technique has been introduced by Neher and
Sakmann [15] and was described in detail by Hamill et al. [21].
It was first used to study channel proteins in biological cells (in
particular the acetylcholine receptor in frog muscle cells). A
small pipette is placed on a cell surface while applying slight
suction. Currents and capacitances are measured by using two
electrodes, one being located in the pipette and the other one
in the cell. The patch pipettes have much smaller diameters
of order 1-10 µm (see Fig. 3, right bottom). The obvious
advantage over the Montal-Mueller technique is that one can
investigate intact cells and on does not have to use organic
4
Figure 4: Comparison of conductance events through bi-
ological membranes containing channel proteins and a
pure lipid membrane in its melting regime. a. Record-
ing of the acetylcholine receptor in frog muscle cells from
the classical paper of Neher & Sakmann [15] (Ringer
solution in the pipette).
b. BLM experiment on a
DOPC:DPPC membrane in 150mM KCl, 60mV in its
transition regime at 19◦C. Figure adapted from [27].
conductance states in an unmodified synthetic lipid film (of
somewhat exotic nature). While the lipid system used does
not allow to draw conclusion with respect to biological mem-
branes, the experiments quite clearly are a proof-of-principle
in so-far as they demonstrate that lipid films alone can generate
phenomena that are typically attribute to proteins, peptide
pores or other additives. The original recording was made
with an analogue chart recorder in which the pen position was
changed by rotation around an axis. This has been digitally
corrected by us (see the original paper for a reference to the
raw data). A can be seen in Fig. 5 (top) a change of the
voltage leads to an overall larger membrane current the shows
quantized steps.
Another important early paper showing this phenomenon
is authored by Antonov and collaborators in 1980 [8]. They
showed that one obtains quantized currents DSPC membranes
one obtains quantized currents of in the pA regime at 100
mV. As in the paper by Yafuso et al.
[7], they found that
current traces may display several steps as shown in the current
histogram in Fig. 5 (bottom). In their paper they state: "We
suggest that these channels could conduct the transmembrane
ionic current in biological membranes without the involvement
of peptides and proteins",
thus making the first statement
indicating that this could be a general phenomenon in biomem-
branes.
Other authors have found the same phenomenon. Fig. 6 (top)
shows results from Yoshikawa et al.
[23]. They described
quantized current fluctuations through DOPC membranes con-
taining some cholesterol that are of order 1 pA (corresponding
to a conductance of 50 pS). The authors found similar results
for membranes in BLMs, on patch pipettes and on membranes
deposited on filter paper. As other authors, they also discuss
the relevance of this finding for the interpretation of currents in
biomembranes and suggest that currents attributed to proteins
can also arise from lipid membrane fluctuations.
Fig. 6 (bottom) shows quantized membrane currents in a
membrane made from POPE:POPC=7:3 mixture (from [28]).
Figure 5: Top: First channel recordings in a lipid mem-
brane made of oxidized cholesterol [7]. The bottom line
indicates the voltage. The dashed lines indicate the volt-
age changes that give rise to large capacitive currents in
the current recording. For constant voltage current fluc-
tuations of the order of 10-20 pA occur. The data were
digitally adapted from the original Figure and numeri-
cally corrected for the distortions caused by the analogue
chart-recorder. Bottom: Quantized currents through syn-
thetic DSPC membranes close to their transition in 1M
KCl at 100mV [8].
Figure 6: Top: Current fluctuations in a mixture of DOPC
and cholesterol (about 9:1) at 20mV in 1 M KCl [23].
Bottom: Conductance fluctuations of a membrane made
of a POPE:POPC=7:3 mixture in 150mM KCl [28].
The currents display a similar conductance as in the above
experiment. The authors discuss the similarity of the lipid
conduction steps with those attributed to proteins channels and
state: "This observation demonstrates the need for caution in
interpreting conductance changes, which occur following ejec-
tion of channel- containing vesicles near a membrane." This
refers to the standard reconstitution procedure of generating
5
Figure 7: Currents through a purely synthetic diphytanoyl
phosphatidylcholine membrane (1M KCl unbuffered, pH
ca. 6.5). Left: Currents recorded at 77mV in differ-
ent scalings. Right: Current histogram showing several
steps. From Kaufmann et al. [29].
synthetic membranes containing proteins.
In particular, the
authors state that currents attributed to proteins could in fact be
due to lipid pores.
[29] showed further interesting results on
Kaufmann et al.
channel appearance in purely synthetic membranes. Their pa-
per appeared as a booklet (available from the website given in
the reference list and from the University Library in Gottingen,
Germany) and gives considerable insight into the physics of
the pore formation process that goes along the line of argument
in section 5.2. A nice example for quantized currents through
membranes is given in Fig. 8 (top) [2]. At least 4 steps in the
current recording through a DOPC:DPPC= 2:1 mixture can be
seen. Each of them corresponds to a conductance of about 300
pS. The experimental temperature of 19◦C is close to the tran-
sition maximum of this lipid mixture. Fig. 8 (bottom) shows
a long trace from a BLM made of a DC15PC:DOPC=95:5
mixture in the phase transition regime (31.5◦C, 1000mV) [30].
The above experiments show that there is quite a range of pos-
sible pore sizes and conductances found for lipid membranes.
We focus here on literature demonstrating quantized current
events through lipid membranes. There is additionally a com-
plete literature on reversible electrical breakdowns of artificial
membrane conductance at higher voltages [31, 32, 33], which
are connected to electroporation. We believe, that lipid channel
formation and electroporation are related phenomena.
4.2 Dependence of channel formation of the
melting transition
As mentioned in section 2 the overall membrane conductivity
is maximum in the transition regime. Consistent with this find-
ing the likelihood of finding lipid channel events is maximum
in the lipid melting regime. Below we show three experiments
that demonstrate this. Boheim et al. [9] reported quantized cur-
rent events in synthetic 1,3-SMPC membrane at the phase tran-
sition temperature of about 29◦C (Fig. 9). The discrete conduc-
tance steps are not well visible in the figure, but were obvious
in the original data trace (Wolfgang Hanke, personal commu-
Top:
Current
in
8:
fluctuations
a
Figure
DOPC:DPPC=2:1 mixture at 19◦C (150mM KCl,
60mV). One can resolve at least 4 quantized steps with a
current amplitude of about 20 pA corresponding to about
to a conductance of about 300pS. From [2]. Bottom:
Current fluctuations in a DC15PC:DOPC=95:5 mixture
at 31.5◦C (1000mV), which is the transition regime of
this mixture. Quantized currents with an amplitude of
15pA can be seen. From [30].
nication). They write: "Apparently, the transition of a bilayer
membrane from the liquid-crystalline to the solid state and vice
versa is associated with spontaneous fluctuations in membrane
conductance." Antonov et al. demonstrated for both DSPC [8]
and DPPC [34] that they found quantized currents in the phase
transition regime while the membrane seemed inert outside of
this regime. Fig.10 (left) shows DPPC membranes above the
transition at 50◦C (top), in the melting regime at 43◦C and be-
Figure 9: Currents through a 1,3-SMPC membrane with
a transition at 29◦ recording during a temperature scan.
At the transition temperature quantized current events
were found (not well resolved in the figure but well de-
scribed in the text). From [9].
6
Figure 10: Currents through DPPC membranes with Tm
around 41◦C. Left: heat capacity profile of DPPC. Right:
Currents through DPPC membranes above (50◦C), close
to (43◦C) and below (35◦C) the transition. Only close
to the transition quantized currents with discrete steps in
the current histogram were seen. From [34].
low the transition at 35◦C. The current fluctuations in the tran-
sition are quantized and are of unusual magnitude (nA regime
indicating very large pores).
Transitions in single lipid membranes occur over a quite nar-
row temperature interval of less than one degree which makes
it difficult to adjust the experimental temperature accurately.
The heat capacity displays a very large amplitude at maximum.
As shown below (section 5.5) that the permeability is closely
related to the heat capacity, and that fluctuations in area are
very large. Thus, in transitions of single lipid membranes they
display a tendency to show very large pores and a pronounced
tendency to rupture. Those black lipid membranes are inher-
ently unstable and break easily.
For the above reason, other authors prefer to work with lipid
mixtures. They display broader melting profiles with smaller
fluctuations at maximum. These membranes are more stable
and due to their wider melting regime not as temperature sen-
sitive. This approach has been taken by [2, 24]. They used a
DOPC:DPPC=3:1 mixture (150mM KCl, 40mV) that displays
a broad melting profile with a maximum around 17◦C.At the
heat capacity maximum many discrete current steps of about
30pA were found (Fig. 11), corresponding to a conductance of
≈ 750 pS. Above the melting temperature at 30 ◦C (in the fluid
phase regime) no discrete current steps were seen.
4.3 Specificity of lipid channels
In their seminal paper from 1980, Antonov et al. [8] compared
quantized currents through membranes made of the zwitteri-
onic DSPC with those through positively charged membranes
of a DSPC analogue in which the P-O− in the head-group
through
BLMs
of
11:
Currents
a
Figure
DOPC:DPPC=3:1 membrane (from [2]). Top: heat
capacity profile with a maximum at 17◦C. The two
dotted lines represent the temperatures at which the
currents in the bottom panel were recorded. Bottom, left:
Current steps could only be observed at the transition
temperature (17◦C, 150mM KCl, 40mV) but not in the
fluid phase regime at 30◦C. Bottom, right: Current
histograms. A baseline current was subtracted.
was replaced by a P-CH3 rendering the membrane positively
charged. They found that the cationic membranes were anion
selective while zwitterionic membranes did not display any
selectivity for charges (unfortunately, no experimental details
were given).
In [34], the same group of authors compared the conductances
of membrane pores for different cations. They recorded the sin-
gle channel currents through DPPC membrane in the presence
of LiCl, NaCl, KCl, RbCl and CsCl. In a voltage regime from
-80 to +80 mV they found a linear current-voltage relationship
from which a conductance gi can be calculated (Ii = gi · U,
where i is the index for the ion species). The details of the
experiment are given in Fig. 12. The authors found that the
value of the conductance decreases in the sequence:
gLi > gN a > gK > gRb > gCs
Antonov and collaborators discussed this sequence in con-
nection with kosmotropes (water structure enhancers) and
chaotropes (water structure breakers) that are closely related to
the Hofmeister series. This series involves a ranking of ions
in respect to their effect on various materials in electrolytes
(e.g., proteins) and is related to the interaction of the ions with
water. The Hofmeister series has been discussed in connec-
7
one might expect the membrane defects form through stochas-
tic processes that favor broad distributions of pore sizes. The
Figure 12: Currents through pores in DPPC membranes
close to the chain melting temperature (43◦). Top: Cur-
rent traces in the presence of LiCl at different voltages
lead to a linear current-voltage relationship. Bottom:
Current-voltage relationships in the presence of LiCl,
NaCl, KCl, RbCl and CsCl yield different conductances.
tion with peptide pore conductances [35]. According to [36]
the sequence from chaotropic to kosmotropic for the above
ions is Li+ > Cs+ > N a+ > K + while it was given as
N a+ > K + > Rb+ > CS+ by M. Chaplin1 consistent with
the conductances in the above experiment.
It is very likely
that such correlations exists, rendering the lipid membrane
moderately ion selective2.
5 Theoretical considerations
5.1 Pore models
The remarkable finding in electrophysiological lipid membrane
experiments is that ion permeation expresses itself as quantized
current events. This all-or-nothing feature is unexpected since
1See: www1.lsbu.ac.uk/water/hofmeist.html
2Selectivity is considered here to be a conduction preference expressed in
different permeation rates rather than an exclusive all-or-nothing conduction
phenomenon.
Figure 13: Two representations for transport across a
membrane. Top: Single file transport. Water molecules
align in a pore.
Center:
Schematic drawing of membrane pores: a hydrophobic
pore. Bottom: a hydrophilic pore. From Glaser et al.
[38].
From Finkelstein [37].
permeability of membranes has been imagined to occur via
three mechanisms:
1. diffusion of solutes through the hydrophobic core of the
membrane
2. transport along water files consisting of single H2O
molecules (Fig.13,left)
3. hydrophobic pore formation (Fig.13, right top)
4. hydrophilic pore formation (Fig.13, right bottom)
Mechanism 1 was proposed for oil-soluble substances like
anesthetics already in the 19th century [39]. However, for
ions or polar molecules as water this is an unlikely mechanism
considering the very low solubility of such substance in the
hydrophobic core membrane core. The free energy of a charge
in a dielectric medium is proportional to the electrostatic po-
tential Ψ = q/0r where q is the charge, 0 is the vacuum
permittivity, is the dielectric constant, and r the distance
from the charge. The dielectric constant in water is about 80
while it is 2-4 in the membrane interior. Thus, the electrostatic
free energy is different by about a factor of 20-40 between
membranes and water. Due to Debye-Huckel ion screening
this effect is even larger in the aqueous electrolyte solutions.
Transport mechanism 2 has been described by Levitt [41] and
Finkelstein [37], see Fig. 13 (top). Finkelstein defines: "By
single-file transport of water one means that water molecules
8
Figure 15: Schematic drawing of the chain of events in
the molecular dynamics simulation of POPC membranes
shown in Fig. 14. After the application of a voltage of
2V across the membrane, first a "prepore" opened that
is both consistent with the water file shown in Fig. 13
(top) and the hydrophobic pore in Fig. 13 (center). Sub-
sequently, a hydrophilic pore with a radius of about 0.5
nm formed. From Bockmann et al. [40].
file formed along which ions could flow. This was called a
'prepore' [40] and occurred after less than 1 ns of simulation
time. After ∼ 50ns a large pore developed long this defect that
had a diameter about 1 nm (Fig. 14, bottom). The chain of
events in the simulation is shown in Fig. 15. First, a water file
(prepore) forms along which ions can flow. This is consistent
with the hydrophobic pore suggested by [38] (Fig. 13, center).
Subsequently, the lipids rearrange and form a hydrophilic pore
(Fig. 13, bottom).
If we assume that this is the likely chain of events then we
would conclude that the single file (prepore) defines a minimum
size of a conducting event in a membrane. The hydrophilic pore
seems to display a stable radius of the order of 0.5 nm. This is
consistent with the magnitude of the currents measured in the
patch-clamp and the Montal-Mueller techniques. Remarkably,
this is also the typical size of the aqueous pores proposed for
channel proteins.
Assuming a pore of radius r, Glaser et al. [38] calculated a free
energy of pore formation of
∆F (r) = 2πγr + πr2(σ + αU 2)
(1)
where γ is the edge tension of the pore, σ the lateral tension
and U the voltage. α is a coefficient related to the capacitance
of the membrane. The free energy therefore consists of three
contributions: the energy of the pore interface, the work that
has to be performed against an external tension σ that is pro-
portional to the pore area, and a correction term containing
voltage that is also proportional to the pore area. One should
assume that the probability of pore formation is related to
exp(−∆F (r)/kT ) and the conductance of one pore is in a
simple manner related to the pore radius r.
The above equation does not yet contain a particular the-
ory for the effect of the phase transition, which is obviously of
major importance as judged by the experimental data shown in
sections 2 and 4. This will be provided below.
Figure 14: Molecular dynamics simulation of a POPC
membrane with a tansmembrane voltage of 2V. Top: Hy-
drophobic prepore in an early stage of the simulation
(less than 1ns). Bottom: Pore with dimensions of the or-
der of 1 nm after about 50ns. Adapted from Bockmann et
al. [40].
within the pore cannot pass or overtake each other." In this
context the term "no-pass transport" has been used. Ions would
be part of the file and can be transported in sequence with the
water molecules. The single file mechanism might explain the
quantized nature of currents because there is a minimum size
for the water file defined by the dimensions of H2O and the
ions.
Pore formation is the most popular mechanism proposed
for membrane transport. Glaser et al.
[38] suggested both
hydrophobic (mechanism 3, Fig.13, center) and hydrophilic
pores (mechanism 4, Fig.13, bottom). The hydrophobic pore is
probably related to the formation of a water file as described
in (2) and does not involve any rearrangement of lipids. The
hydrophilic pore involves a change in the geometry of the lipid
head groups. The elastic free energy of such a pore can be
calculated assuming typical values for the elastic constants.
Stable pores with a diameter of ∼ 1 nm were suggested. On
theoretical grounds it was further proposed that transmem-
brane voltage can stabilize large pores at high voltage [42].
This might be important in connection with the phenomenon
of electroporation [33].
Recent molecular dynamics simulations of membranes ex-
posed to large voltages (∼ 2V) suggest that both single file
and hydrophilic pore formation can occur. Fig. 14 shows a
POPC membrane at two stages during the simulation after a
voltage of ∼ 2V was applied. In Fig. 14 (top) a single water
9
5.2 Entropy,
forces
thermodynamic variables and
In 1910, Albert Einstein wrote a remarkable paper on critical
opalescence [43] in which he considered the phenomenon of
maximum light scattering of liquid mixing close to a critical
point (called "critical opalescence"). This paper contains a
deep reflection of the nature of the second law of thermodynam-
ics and is at the basis of all non-equilibrium thermodynamics
and fluctuation theory. The second law basically states that
the most likely state of a system is most probable. This were
without content if one could not extract more information from
this law without making assumptions on the molecular detail
of a system. Einstein showed in fact that one can make general
statements about the nature of fluctuations. His theory for
critical opalescence is similar to the thermodynamic theory of
the pore formation process outlined below. The application of
this theory for pore formation was already proposed by Nagle
& Scott [4] and Kaufmann et al. [29].
Let us consider the entropy as a function of the extensive
thermodynamic variables α (e.g., energy E, volume V, area A,
charge q, ...) and define the deviations from the equilibrium
value ξi = αi − αi,0 we can expand it into a Taylor series:
(cid:18) ∂S
(cid:19)
(cid:88)
∂ξi
0
i
ξi +
1
2
(cid:18) ∂2S
(cid:88)
(cid:19)
∂ξiξj
0
ij
ξiξj + ...
(2)
S = S0 +
The fact that the entropy is at maximum in an equilibrated sys-
tem implies that the linear terms (∂S/∂ξi)0 are zero and thus
we can approximate the entropy for small fluctuations with
(cid:18) ∂2S
(cid:88)
(cid:19)
∂ξiξj
0
ij
ξiξj ≡ S0 −(cid:88)
ij
S ≈ S0 +
1
2
gijξiξj
(3)
where gij = −0.5(∂2S/∂ξiξj)0 are positive constants.
The expansion of the entropy into a quadratic function of
the intensive thermodynamic variables is at the basis of the
complete discipline of linear non-equilibrium thermodynam-
ics as developed by Onsager [44, 45] and Prigogine [46]. The
probability to find a particular pair of fluctuations, ξi and ξj, is
given by
P (ξi, ξj) = P0 exp
= P0 exp
or for one single fluctuation ξi it is given by
(cid:18) S(ξi, ξj)
k
(cid:19)
(cid:18)
(cid:19)
− gijξiξj
k
(cid:18)
(cid:19)
P (ξi) = P0 exp
− giiξ2
k
i
P0 =(cid:112)gii/π k.
which corresponds to a Gaussian distribution of states with
The thermodynamic average of the product of the fluctuations
Figure 16: The entropy as a harmonic potential and the
corresponding fluctuations of the variable ξ that could be
internal energy, volume, area etc. The slope of the poten-
tial at ξ is the thermodynamic force, while the curvature
of the potential at ξ0 is proportional to the corresponding
susceptibility (cf. eqs. 7, 10 and 12).
ξiξj is given by [47]
(cid:90)
(cid:104)ξiξj(cid:105) =
ξiξjP (ξi, ξj)dξ1dξ2dξ3...dξn = − k
2gij
where (cid:104)...(cid:105) denotes the statistical mean, which implies
(cid:10)ξ2
i
(cid:11) = − k
2gii
(6)
(7)
This corresponds to the mean square amplitude of the fluctu-
ations or the width of the Gaussian distribution. This implies
that the half width of the distribution is related to the curvature
of the entropy potential, a fact that we will make use of below
in order to derive the relaxation times of fluctuations.
The thermodynamic forces are defined as the derivatives of
the entropy with respect to the extensive variable.
= −(cid:88)
j
Xi =
∂S
∂ξi
gijξj
(8)
(4)
(5)
This is completely in analogy to the definition of forces in
classical mechanics (as the derivative of a potential) but more
general since it includes extensive variable as the energy and
charge, but also the particle numbers . Further, the potential
under consideration here is the entropy.
Some of the thermodynamics forces are
1
T
p
T
with the conjugated variable
E (interal energy)
with the conjugated variable
V
(volume)
10
Π
T
Ψ
T
µj
T
with the conjugated variable
A (area)
(9)
with the conjugated variable
q
(charge)
with the conjugated variable
nj
(particle number)
...
where T is the temperature, p is the pressure, Π is the lateral
pressure, Ψ is the electrostatic potential, and µj is the chemical
potential of species j [47, 46]. E, V, A, q and nj correspond to
the fluctuating variables ξi.
The above implies that
(cid:10)ξ2
i
(cid:11) = −k
(cid:18) ∂Xi
(cid:19)−1
∂ξi
= kg−1
ii
(10)
meaning that the fluctuations are related both to the curvature
of the entropy potential of the change of the forces upon the
change of a variable. Eq. 10 has been strictly derived in [47].
5.3 Susceptibilities
The susceptibilities or response functions are functions like the
heat capacity, the compressibility or the capacitance. They are
given by the derivative of an extensive variable with respect to
an intensive variable. E.g., the heat capacity at constant volume
is given by
(cid:18) ∂Q
(cid:19)
(cid:18) ∂E
(cid:19)
=
cV =
(11)
Now, ξE ≡ E − (cid:104)E(cid:105) is one of the quantities that can fluctuate,
while ∂T = −T 2∂(1/T ) is related to the conjugated thermo-
dynamic force, XE = 1/T . In particular,
∂T
∂T
V
V
(cid:18) ∂ξE
(cid:19)
∂XE
V
cV = − 1
T 2
Another way of writing this is
(cid:19)−1
V
= − 1
T 2
(cid:18) ∂XE
(cid:10)E2(cid:11) − (cid:104)E(cid:105)2
∂ξE
cV =
kT 2
(cid:10)ξ2
(cid:11)
E
kT 2
eq.10
= +
(12)
(13)
the heat capacity is proportional
(cid:104)E(cid:105) = (cid:80)
i Ei exp(−Ei/kT )/(cid:80)
to the
This implies that
This relationship could
fluctuations in internal energy.
also be derived in a simpler manner by just differentiating
the statistical mean of the internal energy that is given by
i exp(−Ei/kT ). Thus, the
fluctuation relations would be strictly true even if the assump-
tion of a harmonic entropy potential is not made. Similar
relations can be written for the heat capacity at constant
pressure
cp =
(cid:10)H 2(cid:11) − (cid:104)H(cid:105)2
(cid:19)
(cid:18) ∂V
kT 2
=
∂p
T
(cid:10)V 2(cid:11) − (cid:104)V (cid:105)2
V kT
T = − 1
κV
V
the isothermal volume compressibility
(14)
(15)
11
and the isothermal area compressibility
(cid:18) ∂A
(cid:19)
=
∂Π
T
(cid:10)A2(cid:11) − (cid:104)A(cid:105)2
A kT
T = − 1
κA
A
(16)
Eqs. 13 to 16 are linked to the fluctuation theory (see, e.g.,
[47]). The principal underlying idea is that small forces gener-
ated thermal motion generate local fluctuations in the extensive
variables. The coupling of the susceptibilities to the fluctua-
tions is important and will play an crucial role below.
5.4 Coupling of compressibilities with the heat
capacity
It has been argued by [48, 49] that for transitions in membranes
one finds that
∆V (T ) = γV · ∆H(T )
(17)
where ∆V (T ) is the temperature dependent volume change in
the transition regime, and ∆H(T ) is the corresponding change
in enthalpy. This means that volume and enthalpy change in a
proportional manner with a proportional constant γV = 7.8 ·
10−10m3/J for many lipid systems including biological mem-
branes [49]. More indirectly, it has been argued that a similar
relation is also true for area changes in melting transitions of
membranes:
∆A(T ) = γA · ∆H(T )
(18)
with γA = 0.89m2/J [48]. The two above relation together
with eqs. 15 and 16 imply that the fluctuations of enthalpy are
proportional to the fluctuations in volume or area (i.e., enthalpy,
volume and area are no independent variables), and we obtain:
∆κV
T =
γ2
V T
V
∆cp
and
∆κA
T =
γ2
AT
A
∆cp
(19)
This means that both volume and area compressibility are pro-
portional to the heat capacity changes in membrane transitions.
In particular it implies that membranes are very soft in their
transitions. This change in compressibility can easily of order
10-fold at the transition maximum. This has profound conse-
quences for the pore formation process.
5.5 Permeability
A couple of authors have discussed area fluctuations as the ori-
gin of membrane pores (or lipid ion channels), e.g. [3, 50, 6,
2, 51] but in particular Nagle & Scott [4] and Kaufmann [29]
(much of the line of argument in sections 5.2 and 5.3 was given
in the latter reference and applied to membrane pores).
There are in particular two views:
1. In the transition one finds phase coexistence and the like-
lihood of finding a membrane pore is at maximum at the
domains boundary. Therefore, the permeability is pro-
portional to the length of the domain interfaces. This
view has been taken by Papahadjpoulos et al.
[3] and
Mouritsen and coworkers [50, 52].
2. In order for a membrane pore to be created work has to
be performed that is proportional to the lateral compress-
ibility [4, 2, 51].
These two views are conceptually similar but not identical.
Clearly, the domains occur in temperature regimes with larger
area fluctuations. More precisely, the fluctuations in domain
sizes are responsible for both the large heat capacity and the
large compressibility. Domain sizes in lipid transitions depend
on the cooperativity of the melting process. In computer sim-
ulations, cooperativity is usually modeled by an unlike nearest
neighbor interaction that is effectively a free energy penalty
for the creation of a domain interface. For this reason, more
cooperative transitions with narrower and more pronounced
heat capacity peaks display larger domains at the melting
temperature (50% of the lipid matrix is in a liquid state) and
consequently a smaller overall domain interface than less
cooperative transitions. Thus, large fluctuations imply that
domains are large and that the total domain interface is smaller.
The data shown, e.g., in Fig. 2 (right) indicate, however, that
the permeability is about proportional to the heat capacity.
This favors the second view. The discrepancy between the
two views would disappear if one would not consider domain
interface is one-dimensional lines but rather as narrow areas of
large fluctuations. However, under these circumstances, one
would have to take both length and thickness of the interfacial
region into account. For these reasons we follow the argument
that the overall fluctuations are responsible for the increase
in permeability. Domain interfaces are discussed again in
section 8.1 and it is shown that the fluctuations within different
domains may also be different.
Nagle and Scott [4] calculated the free energy (the work)
necessary for the formation of a pore of area ∆A in a membrane
with total area to be
∆F =
1
2
(20)
(∆A)2
1
κA
T A
with the lateral compressibility κA
T of the membrane. This cor-
responds to Hooke's law for a mechanical spring. The likeli-
hood to find a pore of size ∆A is given by
− ∆F
kT
P (∆A) = P0 exp
(cid:18)
(cid:19)
(21)
which implies that the likelihood for a pore is the same when
T is the same. This means one finds pores of twice
(∆A)2/κA
the area if the compressibility is four times larger, or that the
likely pore size scales with
(∆A)2 ∝ κA
T
(22)
As a next step it was assumed that the permeability P of the
membrane is related to the size of the defects such that one can
make a series expansion of the permeability as a function of
area change ∆A:
P = P0 + C1∆A + C2(∆A)2 + ....
(23)
12
where P0 is the permeability of the pure phase in the absence of
fluctuations. In the transition one expects fluctuations ∆A that
are both positive and negative and the linear term is neglected,
yielding
P = P0 + C2(∆A)2 eq.22
= P0 + C(cid:48)
2κA
T
(24)
This equation reflects that one expects larger pore areas ∆A or
a larger number of pores if the compressibility is higher. Nagle
& Scott [4] argued that one obtains the same law if one takes
into account that one finds distributions of pore sizes.
In eq. 19 we have derived that for lipid membranes in tran-
sitions the compressibility changes are proportional to the ex-
cess heat capacity. Therefore, we finally obtain for the perme-
ability P :
P = P0 + α∆cP
(25)
where α is a constant. To summarize, one expects that the mean
pore size and the pore number is larger when the compressibil-
ity is higher. This leads to a higher permeability. For the above
calculation, it does not make a difference if a pore has twice the
area or if one simultaneously finds two pores. Obviously, in the
above derivation the free energy of the interface between pore
wall in water has been neglected. This can only be justified by
experiment. The data in Fig. 2 (right) which yield a satisfactory
proportionality between heat capacity and permeability suggest
that in fact the free energy of a pore is dominated by the area
work as assumed in eqs.24 and 25. This neglects the edge term
of the pore which was contained in eq. 1. The effect of voltage,
that was heuristically included in eq. 1 is also not considered
here because a satisfying description of the effect of voltage on
the melting transition is yet missing, even though experimental
evidence suggests that this effect must be significant (see also
section 6.4).
6
Influence of changes of variables on
the lipid channels
From the above derivations it is clear the changes in all inten-
sive variables should have a predictable influence on the perme-
ability when their effect on the melting transition is known. Ac-
cording to the list given in eq. 9 this includes temperature, hy-
drostatic pressure, lateral pressure, electrostatic potential (volt-
age) and the chemical potentials (e.g., protein, proton or cal-
cium concentrations). In this paragraph it will be shown that
all of these dependencies exist. This is especially so if the lipid
membrane is close to a melting transition (or any other kind of
transition). It was shown by [20] (findings are also reviewed
in [13]) that biomembranes are in fact often found close to a
melting transition in their membranes under physiological con-
ditions (typically about 10-15◦ above the heat capacity maxi-
mum of a relatively broad transition). Therefore, any variable
that potentially changes the transition will display an influence
on the occurrence of lipid ion channel events. That this is in
fact the case is shown below for a number of thermodynamic
variables.
6.1 Temperature
The previous section justified why the formation of lipid chan-
nels is most likely in the melting transition of lipid membranes.
This implies that temperature must be an important variable.
Let us consider the following very simple mass action scheme
for a transition at constant pressure
K =
[F ]
[G]
= exp
− ∆G
RT
= exp
− ∆H − T ∆S
RT
(26)
(cid:18)
(cid:19)
(cid:18)
(cid:19)
where [F ] is the concentration (or fraction) of fluid lipids and
[G] is the concentration of gel lipids, respectively. In the fol-
lowing we will assume that the state of the membrane and its
heat capacity depends on ∆G. For ∆G > 0 we find a larger
fraction of gel phase, while for ∆G < 0 the membrane is
mostly in its fluid phase. The melting temperature, Tm, shall
be defined as the temperature where [G] = [F ], or K = 1 and
∆H − Tm∆S = 0 →
∆S =
∆H
Tm
→ ∆G = ∆H
(cid:18) Tm − T
(cid:19)
(27)
Tm
Thus, obviously a change in temperature will lead to a change
of the free energy difference between gel and fluid phase.
Since at the melting point, Tm, ∆G is zero, the fluctuations
are maximum because the free energy necessary to move the
system from one state into the other one is zero, and thermal
noise is sufficient to do so. Therefore, the likelihood to form
pores and the permeability must be strongly influences by
temperature. This was shown for the macroscopic permeability
experiments in section 2 [3, 6, 5, 2] and for the lipid channel
formation in section 4.2 [34, 2].
The effect of temperature could be expressed as being a
'thermo-sensitivity' of channel formation in pure lipid mem-
branes.
6.2 Pressure and lateral tension
In the following, we will express the state of the system by the
Gibbs free energy difference between fluid and gel state. We
will show that it depends not only on temperature but also on
all the other intensive thermodynamic variables.
The enthalpy is a function of pressure: H = E + pV . If
∆H0 is the enthalpy difference between fluid and gel phase at
a pressure of 1 bar, ∆V the respective volume difference, and
∆p the pressure difference with respect to 1 bar, we obtain
Tm(∆p) =
∆H
∆S
=
∆H0 + ∆p∆V
∆S
13
Figure 17: Channel events in a membrane made of soy-
bean phosphatidylcholine in 1M unbuffered NaCl, glass
pipette tip diameter ca. 5µm, at 500 mV. The top trace (A)
recorded before application of suction does not show any
current fluctuations. B. The bottom traces were recorded
with suction applied. The change in membrane area was
monitored by recording the membrane capacitance that
increased from 260 to 510 pF. From [29].
=
∆H0(1 + γV ∆p)
∆S
= Tm(1 + γV ∆p)
(28)
or
∆Tm(∆p) = γV Tm∆p
(29)
where γV = 7.8· 10−19m3/J is the parameter defined in eq. 17.
Similarly, the shift of the melting transition by lateral pressure
(or tension) is
∆Tm,(∆Π) = γATm∆Π
(30)
with γA = 0.89 m2/J (see eq. 18).
It has been shown by
Kaufmann et al.
[29] that the application of suction on a
patch pipette can generate channel events. Suction changes the
lateral tension/pressure of the lipid membrane and thus has an
influence on the state of the lipid membrane. Fig. 17 shows
suction induced channel formation in soybean phosphatidyl-
choline in 1M NaCl at 500mV.
The influence of pressure on channel appearance could be
called a 'mechano-sensitivity' of channel formation in pure
lipid membranes.
Figure 18: Effect of the anesthetic octanol on the melting transition of a DOPC-DPPC=2:1 mixture in 150mM
KCl at 19◦C. Left: Heat capacity profiles in the presence of increasing quantities of octanol demonstrate the shift
of the transition caused by the anesthetic drug. The vertical dashed line indicates the experimental temperature.
Center: Membrane current recorded at 210 mV demonstrates that octanol both decreases the frequency of current
events and the amplitude until currents are completely 'blocked'. Right: Current histograms of the center traces.
6.3 Anesthetics, neurotransmitters and other
drugs
Anesthetics can lower melting temperature by ∆Tm,xA due to
a very simple effect called melting point depression [53]. It is
described by
∆Tm,xA =
xA
(31)
R T 2
m
∆H
where xA is the molar fraction of anesthetics dissolved in the
membrane, ∆H is the melting enthalpy of the membrane, and
Tm is the melting temperature. To arrive at this law one has
to assume that the anesthetics molecules dissolve ideally in the
fluid lipid membrane, while they are insoluble in the gel mem-
brane. The success of eq. 31 in describing the effect of anes-
thetics on melting transition [53] does not only justifies this
assumption but also explains the famous Meyer-Overton rule
that states that the effectiveness of an anesthetic is proportional
to its solubility in oil (the membrane core), independent of its
chemical composition [54, 55, 56].
Fig. 18 shows the effect of the anesthetic octanol on mem-
branes made of a DOPC-DPPC mixture. In the absence of oc-
tanol, the experiment is performed at the heat capacity max-
imum where many channel events can be found.
Increasing
amounts of octanol shift the transition maximum towards lower
temperatures and thereby lower the heat capacity at the ex-
perimental temperature. As a consequence, the frequency and
amplitude of the current events decreases until channel events
completely disappear. This effect is also discussed in detail in
[2, 51]. If the membrane is recorded at a temperature below that
transition, anesthetics can drastically increase the membrane
permeability because they move it into the transition regime
[2].
Sabra et al. [5] showed that the macroscopic permeability of
DMPC membranes for Co2+ was altered by the insecticide lin-
14
dane in agreement with the influence of the drug on the heat
capacity profile. Other drugs as neurotransmitters also have the
potential to shift melting transitions and to alter the fluctuations
in membranes. This was for example shown for serotonine [57]
that has a quite significant influence on melting profiles. Thus,
neurotransmitters also should display the potential to influence
lipid channel events considerably. More so, every drug that al-
ters the melting behavior of a membrane can change the appear-
ance of lipid channels. This includes proteins and peptides that
can alter lipid channel behavior without being channels them-
selves, as discussed in section 8.2.
The possibility to "block" or induce lipid channel events in
protein-free membranes by drugs could be considered as an
agonist or antagonists effect, or "channel-gating". It does not
require specific binding to any macromolecule but rather repre-
sents an unspecific physicochemical process.
6.4 Voltage, pH and calcium
About 10% of the lipids in most biomembranes carry negative
charges in their head groups. Charged membranes display an
electrostatic potential that depends on the charge density. At
typical conditions of biological membranes, i.e., low charge
density (10% of the lipids) and high ionic strength (≈150mM)
the electrostatic surface potential of a membrane Ψ0 can be ap-
proximated by the low potential approximation that is given by
Ψ0 =
1
0κ
σ ∝ σ√
c
(32)
where 0 is the vacuum permittivity, is the dielectric constant
of water (≈ 80) and σ = q/A is the charge density. κ is the De-
c with c being the ionic
bye constant that is proportional to
strength. Consequently, the larger the ionic strength and the
smaller the surface charge density, the smaller the surface po-
√
tential. From this the electrostatic free energy can be calcu-
lated:
(cid:90) σ
Fel =
Ψ0dq ∝ q2
√
A
σ(cid:48)=0
c
(33)
i.e., the potential is quadratic in the total surface charge and in-
a lowering of the phase transition temperature by
∆Tm,el =
∆Fel,L
∆S
(34)
where ∆Fel,L < 0 is the electrostatic free energy difference
between fluid and gel lipids. The mathematical expression
for ∆Fel,L (not given here) has been derived in [61] and been
reviewed in [13]. The above equation implies that the change
in lipid state from gel to fluid will lower the free energy of
the membrane because the membrane area increases and the
potential decreases. Further, by reversing the argument it is
obvious that application of voltage across a membrane must
change the lipid state.
In particular, charged and uncharged
membranes should respond differently to voltage. Since nega-
tive lipid head groups can be protonated upon lowering the pH,
the membrane state must be pH dependent.
Voltage: The above argument showed that the lipid state
Figure 19: Currents through a membrane made of
DOPC:DPPC=2:1 mixture at 19◦ and 150mM KCl
recorded at different voltages. One finds a linear current-
voltage relationship for Vm <150 mV that becomes
non-linear at higher voltages indicating that the voltage
alters the state of the membrane. From [51].
versely proportional to the membrane area. Since biological
membranes are asymmetrically charged (typically with most
negative lipids being on the cytosolic side of the cell membrane
[58, 59]) one expects a transmembrane voltage of order 50mV
[60]. The molecular area of the gel state of a lipid membrane is
about 25% smaller than that of a fluid lipid [48]. This implies
that the electrostatic free energy of a gel membrane is larger
than that of a fluid lipid membrane because the charge density
is larger in the gel state. The decrease of the electrostatic free
energy contribution upon membrane melting, ∆Fel,L, leads to
15
Figure 20: Effect of charge on the transition temperature
of dimyristoyl methylphosphatidic acid membranes as a
function of pH for four different ionic strengths (adapted
from [61]). The transition midpoint changes by about 14
K towards higher temperature upon protonation of the
membrane. The pKA of the membrane is at higher pH
values when the salt concentration is lower.
should depend on transmembrane voltage. An indication
for this is shown in Fig.19 that displays the transmembrane
currents as a function of voltage, and the corresponding
current-voltage relationship of the channels. It can be seen that
the I-V curve is linear in a regime from about -150 to +150 mV
(i.e., conductance is constant), while the I-V relation becomes
non-linear at Vm > 150 mV. That voltage can reversibly in-
duce channel events was reported by [29]. Antonov et al. [62]
showed that voltage increases the melting transition in charged
lipid membranes without significantly altering the transition
temperature of zwitterionic lipids. However, in our daily lab
routing we reproducibly find that zwitterionic membranes not
displaying channels at low voltage show such events upon
increasing the voltage (in agreement with eq. 1). Therefore,
we assume that voltage also has an effect on zwitterionic mem-
branes. We further suspect that lipid ion channel formation is
closely linked to the phenomenon of electroporation (cf. [33]).
It should be added zwitterionic membranes are piezoelectric
due to the electrical dipoles of their head groups [63]. This was
also shown in monolayer experiments demonstrating different
dipole potentials in the gel and the fluid phases of the mem-
branes and an effect of voltage on the lipid state [64].
The influence of voltage could be called 'voltage-sensitivity'
or 'voltage-gating'.
Calcium: Due to its two positive charges, calcium also dis-
plays a pronounced effect on lipid membranes. It binds to both,
negatively charged and somewhat weaker to zwitterionic mem-
branes. This is likely to be associated with the electrostatics
of the membranes. The two calcium charges probably cross-
link the phosphate groups of two adjacent lipids. Typically,
calcium increases melting temperatures. Depending on state of
the membrane (e.g., whether found below or above a transition)
calcium may therefore induce or inhibit lipid ion channels. This
is shown in Fig. 22. Gogelein & Koepsell [65] showed that
channel events in brain phosphatidylserine can be "blocked" or
inhibited by calcium (Fig. 22, top, while Antonov et al. [12]
showed the reverse effect for DPPC membranes above the tran-
sition (Fig. 22, bottom). The effect of calcium corresponds to
Transmembrane currents through soy-
Figure 21:
bean phosphatidylcholine membranes in an unbuffered
medium with 1M KCl. Traces are shown with different
proton concentrations (pH range from 6.5 (trace a) to 1
(traces f and g)) at different positive and negative volt-
ages. Transmembrane currents are induced at low pH
close to the pKA of the membrane. From [10].
Protons and pH: Let us consider a membrane that is
negatively charged at neutral pH. Upon lowering the pH the
phosphate groups will be protonated. This case has been
treated in much detail by Trauble and collaborators [61]. They
provided the electrostatic theory briefly outlined above and
showed how the melting temperature depends on pH and ionic
strength. Upon complete protonation the transition midpoints
increases by 10-20 degrees. This is shown in Fig. 20 for
dimyristoyl methylphosphatidic acid (DMMPA) membranes
[61]. This is shown in Fig. 20. Similarly, zwitterionic lipids
my become become positively charged at
low pH values
around1-2. Their transition midpoint will decrease due to the
electrostatic contribution of the lipid charges. From this it must
be followed that the occurrence of lipid ion channels should
be pH-sensitive. This has in fact been shown by Kaufmann
and Silman [10], see Fig. 21. They found that channels can
be induced by protons in the pH regime around 1 that is close
to the pKA of the zwitterionic soybean phosphatidylcholine
membrane.
16
Influence of calcium on the occurrence of
Figure 22:
lipid channels. Top: Currents through bovine brain phos-
phatidylserine : cholesterol=2:1 (w/w) membranes in
200mM cyclamate and pH 7.4. Addition of 8mM CaCl2
inhibited the currents. From [65]. Bottom: Generation
of quantized currents in DPPC membranes by calcium
at T=64◦C (the melting temperature of DPPC is 41◦C).
From [12].
a 'calcium-control' or 'calcium-gating' of the lipid channels.
As for the case of drugs, no specific binding site at a particu-
lar macromolecule is involved. One should bear in mind that
both pH and calcium effects are related to the surface poten-
tial and thus to the voltage effects. This includes the adaptation
of lipid composition reported for many bacteria under various
temperature, pressure or solvent conditions (see below).
6.5 Other variables
Everything that potentially changes the state of the lipid mem-
brane will display the potential to generate or inhibit lipid ion
channels. We just list a few of these:
• membrane curvature, because one of the work terms in
the internal energy differential is related to curvature, and
curvature has an influence on the membrane phase be-
havior [66].
• dipolar interactions, because the work to orient a dipole
(e.g., the lipid head group) in a field is part of the dif-
ferential of the internal energy. It has been shown that
electrostatic fields can alter the phase behavior of zwitte-
rionic monolayers [64].
• hydrolysis of membranes by enzymes, because hydroly-
sis products change the phase behavior of the membranes
[67]. Any change in membrane composition may have a
similar effect.
• membrane proteins. This case is discussed in section 8.2
• ...
6.6 Free energy of channel formation
Let us consider a membrane in a particular state that is not
within the melting transition. This membrane displays a very
small likelihood to display lipid ion channels. In order to in-
crease the channel probability, the membrane has to be brought
into the chain melting regime, which can be done be altering
the various parameters described above. Heimburg & Jackson
[53] showed that the free energy that is necessary to bring the
membrane into a regime where channels occur (i.e., the free
energy difference between the fluid and the gel phase) is given
by:
∆G(T, xA, ∆p, ...) =
Tm − T
Tm(cid:124) (cid:123)(cid:122) (cid:125)
temperature
∆H
− RT
∆H
(cid:124) (cid:123)(cid:122) (cid:125)
anesthetics
(cid:124)
(cid:123)(cid:122)
T
Tm
(cid:125)
pressure
xA
+ γV ∆p
+...
(35)
For each of the thermodynamic variables, there is one term in-
fluencing the free energy difference. One could add terms for
the electrostatic potential, for pH, calcium concentration, cur-
vature, etc., which has not been done here. The content of this
equation is: The further away a membrane is from its melting
transition the more free energy has to be provided to generate
channels. Further, for a given value of the free energy differ-
ence the likelihood of channel formation is constant, but the
values of the individual variables can change. The effect of
anesthetics can be compensated by hydrostatic pressure, pH,
or temperature change. These effects have been documented
for anesthesia [68, 69, 53] yielding the well-known effects of
pressure-reversal of anesthesia and the impossibility to anes-
thetize inflamed tissue with normal anesthetic doses.
17
7 Channel lifetimes
It has been found in many experiments that relaxation time
scales are closely coupled to the melting process in membrane
[70, 71, 72, 73, 74, 75, 76, 57]. The fluctuation-dissipation the-
orem reveals a fundamental coupling between the magnitude of
fluctuations (expressed in heat capacity, compressibilities, ca-
pacitance, ...) and the the relaxation time scales (indicating how
fast a system relaxes to equilibrium after a perturbation), which
are identical to the fluctuation time scales. Since we consider
here the lipid channel formation process basically as the con-
sequence of a local fluctuation in state, it is obvious that open
and closed life times of a channel must be intimately related to
the fluctuation time scale. Basically, the fluctuation-dissipation
theorem states that at regimes of high compressibility and high
heat capacity, fluctuations are slow. Thus, channel life times
are expected to be long in the phase transition regime when
their likelihood of formation is large.
7.1 Theoretical considerations
We now return to the theoretical treatment of fluctuations given
in sections 5.2 and 5.3. On the basis of this description the fluc-
tuation life times of membranes can be obtained [76, 57]. Lars
Onsager [44] treated time dependent phenomena such that he
assumed a proportional relation between fluxes (rates of change
of a variable) with the thermodynamic forces, an assumption
corresponding to the equation of motion in viscous fluids where
the influence of inertia is small. Following this approach, the
flux of heat (enthalpy) of the membrane after a perturbation
back to equilibrium (which could consist of the thermal colli-
sions related to Brownian motion) is given by
dξH
dt
= L · XH
(36)
where XH is the thermodynamic force related to enthalpy dif-
ferences (cf. section 5.2), given by ∂S/∂H with the thermo-
dynamic variable ξH = H − (cid:104)H(cid:105). L is a material constant
that we assume to be temperature independent. This equation
is a simple version of Onsager's phenomenological equations if
only one independent thermodynamic variable is present.
= −gHH · (H − (cid:104)H(cid:105))
(37)
According to eqs. 10, 12 and 13,
XH =
∂S
∂H
(cid:10)H 2(cid:11) − (cid:104)H(cid:105)2
cp =
kT 2
=
1
T 2gHH
or
gHH =
1
T 2cp
(38)
This leads to
d(H − (cid:104)H(cid:105))
dt
= − L
T 2cp
(H − (cid:104)H(cid:105))
(39)
with the solution
(H − (cid:104)H(cid:105))(t) = (H − (cid:104)H(cid:105))0 · exp
≡ (H − (cid:104)H(cid:105))0 · exp
introducing the relaxation time
τ =
T 2cp
L
(cid:18)
(cid:18)
− t
τ
− L
T 2cp
(cid:19)
(cid:19)
t
(40)
(41)
This equation implies that the relaxation (or fluctuation) time
scale is large when the heat capacity is high. If we consider
the channel formation as a fluctuation, we have to conclude
that the typical time scale of channel opening and closing is
proportional to the heat capacity. In particular, in the transition
where the likelihood of lipid ion channel formation is high, they
should also display long open times.
It has been found that for pure lipid membranes the phe-
nomenological coefficient L is of order 6-20·108 J K/mol sec.
For a DMPC multilayer one finds a relaxation time of about 30
seconds at the maximum. Transitions in biological membranes
are much broader and the heat capacity maximum displays
much smaller values. Assuming the same phenomenological
coefficient as for DMPC membranes, one estimates about 100
milliseconds at the cp maximum of bovine lung surfactant or
E.coli membranes. Since biological membranes display a cp
maximum some degrees below body temperature, the expected
open times at physiological temperature is expected to be of
the order of a few milliseconds. Interestingly, this time scale is
just of the same order as the dwell times of quantized current
events in section 4.
Summarizing: While the channel open times in artificial
membranes at the cp maximum may be very long (of order sec-
onds), they are expected to be of millisecond order for biomem-
branes at physiological temperature. This time scale, expected
for lipid channels, is interestingly also exactly that found for
protein channels. This is unlikely to be accidental.
7.2 Experiments
It has been shown by [30] that channel life times are in fact
much longer in the transition of a lipid mixture that displays
a transition half width comparable to that of biomembranes.
This is shown in Fig.23. Both heat capacity profile and mean
conductance of a lipid mixture of DC15PC and DOPC (95:5)
display maxima at about 30◦C (Fig.23, top). The typical open
time of the lipid channels is of order 20 milliseconds, while it
is rather 3ms above the transition. If one compares the time
scales of the current fluctuations in the phase transition regime
and in the fluid phase one finds that they are about one order
of magnitude different (Fig.23, bottom), in approximate agree-
ment with the above calculation (eq. 41). The timescale in
the fluid phase follows a single-exponential decay while in the
18
Figure 23: Top: Relative conductance of a D15PC-
DOPC (95/5) mixture as a function of temperature at Vm
= 90mV. The insert shows the melting profile of the lipid
mixture. Adapted from [30]. Bottom: Logarithmic plot of
the timescales of the current fluctuation in the fluid phase
(circles) and the phase transition regime (triangles). The
average lifetimes center around 3ms in the fluid and 20ms
in the phase transition, which is in good agreement with
our theoretical prediction. From [30].
transition it can only be approximated by a double-exponential
function. Gallaher et al.
[77] provided an analysis showing
that the current fluctuation time scales in the transition regime
obey a power law rather than single exponential behavior as
suggested by the derivation above. The authors rationalized
this by assuming that the conduction pores may exist in dif-
ferent states with different gel-fluid environments. Figs. 13
and 14 also suggest that there might be more than one possible
pore geometry with similar pore dimension.
Summarizing, it seems as if one can understand the time scale
of lipid ion channel opening and closing with simple non-
equilibrium thermodynamics (fluctuation-dissipation). This
means that heat capacity, phase behavior and time scales are
different aspects of the same physics. Whenever heat capacities
and permeabilities are high, time scales are larger. However,
while the overall change in time scales is consistent with the
above derivation, multi-exponential or power law behavior is
not contained and requires more sophisticated analysis.
Heat capacity profiles can be influenced by anesthetics, neu-
rotransmitters and by membrane peptides. Interestingly, it has
been shown by [57] that those cp changes are accompanied
by equivalent changes in the relaxation time scales, such that
those molecules cause a predictable change of the relaxation
time scales. The role of proteins is outlined in more detail in
the following section.
8
Interfaces: Domains and proteins
Fluctuations are large in regions where the free energy of gel
and fluid lipids are similar. Therefore, they are maximum at the
melting transition. As shown above, many variables can be ad-
justed (pressure, temperature, ...) such that the membrane sys-
tem is situated in this regime. However - such fluctuations must
not be the same everywhere in the membrane if the membrane
is heterogenous. Biological membranes are complex mixtures
of lipids with different melting points and different hydropho-
bic lengths, and many other molecules including cholesterol
and proteins. These membranes display phase separation and
domain formation phenomena. Therefore, one must expect that
also the fluctuations display different magnitude in different
membrane regions. It is especially interesting to consider in-
terfaces between fluid and gel domains.
The magnitude of fluctuations depends on the length scale over
which they are monitored. One can consider fluctuations of
the whole membrane, which enter into the heat capacity or the
macroscopic compressibility. If one considers the fluctuations
in windows of smaller size, e.g. on the length scale of indi-
vidual domains or on molecular scale, one obtains local heat
capacities and compressibilities. There are different ways of
probing the local fluctuations, e.g.,
• experimentally: One can investigate the softness of a
membrane by AFM. One obtains the compressibility on
the length scale of an AFM-tip (around 10nm). This has
been done for peptide containing membranes in [78].
• theoretically by simulations: In Monte Carlo simulation,
one can directly obtain the fluctuation on arbitrary scales.
Such simulations typically use experimental observables
such as melting enthalpy and entropy, and the shape of
the melting profile in order to determine the simulation
parameters (e.g., [79, 80, 81, 82, 83]). In principle, such
information is also contained in molecular dynamics
(MD) simulations. However, to our knowledge MD has
not been used yet to obtain such information.
8.1 Domain interfaces
In Fig. 24 domain formation of lipid mixtures is shown. Confo-
cal microscopy images of large unilamellar vesicles display do-
mains of gel and fluid state (Fig. 24, top). Such domain forma-
tion can be described by lattice Monte Carlo simulations (Fig.
24, bottom left). Details are, e.g., given in [85]. For the purpose
19
Figure 24: Top: Confocal microscopy image of a DLPC-
DPPC (1:2) mixture in its melting regime at 27◦C. Dark
regions represent gel domains, bright regions are fluid
[84]. Bottom, left: Snapshot of a Monte Carlo simulation
of a DMPC-DSPC (1:1) mixture in the melting regime
(36.9◦C). Dark shades correspond to gel domains while
bright shades represent fluid domains. Bottom, right: Lo-
cal fluctuations of the same matrix. Bright shades corre-
spond to large fluctuations while dark shades correspond
to small fluctuations. The magnitude of the fluctuations
is related to the membrane permeability and the likeli-
hood of lipid channel formation. For the example shown,
fluctuations are larger in the fluid regions than in the gel
regions, but are maximum at the interface between the do-
mains. These regions are those of maximum permeability.
From [85].
of this review only the results are important. The fluctuations
of single sites in this matrix (local fluctuations) are shown in
Fig. 24 (bottom right). It is found that under the conditions
indicated in the figure legend the gel domains display much
smaller local fluctuations than the fluid domains. The largest
fluctuations, however, are seen at the domain interface. Thus,
while the domain interfaces are the regions of largest perme-
ability, also the domains themselves contribute to permeation
(in Fig.24 (bottom right) especially the fluid domains). This
sheds further light on the two possible explanations for the lo-
cation of permeation pores discussed in section 5.5.
is a variation of a picture from Mouritsen & Bloom used to
describe their mattress model [88]. If a protein with large hy-
drophobic length is penetrating a fluid membrane with smaller
thickness, the lipids at the interface of the protein will be close
to the melting regime, which is due to wetting of the hydropho-
bic core of the protein (Fig. 27, top). Thus, at this interface fluc-
tuations are large and the likelihood of lipid channel formation
is maximum. Similarly, if the protein has a short hydrophobic
8.2 Proteins as membrane perturbations
The gel-fluid domain interface is not the only region of large
fluctuations. Macromolecules as proteins can also create in-
terfaces with lipids that influence their phase behavior. Two
examples are given in Fig. 25, showing the influence of cy-
tochrome b5 and band 3 protein of erythrocytes on the melting
behavior of DPPC and DMPC membranes. While cyt. b5
lowers the melting point, band 3 protein is increasing it.
It
has been shown in [83] that this can be understood if one
assumes that the first protein mixes well with fluid membranes
but not with gel membranes, while the second protein mixes
better with gel and not so well with fluid regions. Mouritsen &
Bloom [88] proposed that this is due to hydrophobic matching
that compares the length of the hydrophobic core of the integral
protein with the thickness of the membrane. Good miscibility
suggests that the hydrophobic core of the protein has a similar
length then the surrounding membrane. Bad miscibility indi-
cates hydrophobic mismatch.
Gramicidin A is a small dimeric channel peptide. As shown
in Fig. 26 (top) it is significantly shorter than a gel lipid of
DPPC. Thus, one expects unfavorable hydrophobic matching
in the gel phase of DPPC. Fig. 26 (botto left) shows the distri-
bution of gel and fluid state lipids and the proteins as obtained
from a Monte Carlo simulation that was based on heat capac-
ity analysis [83]. The simulation was performed slightly below
the melting temperature of DPPC. While most of the matrix is
found in its gel state with some remaining fluid domains, the
proteins aggregate into smaller clusters. Fig. 26 (bottom right)
shows the corresponding local fluctuations. One finds that the
fluctuations are maximum at the interface of the protein clus-
ters. This indicates that the protein interface also forms a mem-
brane region with larger likelihood of lipid channel formation.
The results from the simulation are summarized in Fig. 27 that
Figure 26: Simulation of the antibiotic peptide grami-
cidin A in a DPPC membrane. Top: Gramicidin A is
very short as compared to the gel lipids of DPPC and
displays an unfavorable hydrophobic matching. Bottom,
left: Monte Carlo snapshot of a simulation of gramicidin
A in a DPPC membrane. Slightly below the melting tran-
sition one finds the membrane mostly in the gel state (gel
lipids - dark grey) with some remaining fluid domains
present (fluid lipids - light grey). Due to the unfavorable
hydrophobic matching, the proteins aggregate in the gel
phase into clusters (black dots - gramicidin molecules).
Bottom, right: For the same snapshot the magnitude of
the fluctuations is shown locally. Dark grey shades corre-
spond to small fluctuations while bright shades represent
large fluctuations with a high likelihood of pore forma-
tion.
Figure 25: Influence of proteins on the melting behav-
ior of lipid membranes. Left: Cytochrome b5 lowers the
melting point (from [13] adapted from [86]). Right: Band
3 protein of erythrocytes increases the melting point (from
[13] adapted from [87]).
20
a striking fact that one finds such events in membranes that do
not contain proteins. The currents display typical amplitudes in
the 10pA regime at voltages around 100mV. Typical life times
are of order 10 milliseconds but can be much longer in some
experiments. These currents can be inhibited by anesthetics [2]
but most likely by everything that shifts melting transitions, in-
cluding peptides and neurotransmitters. They respond to pres-
sure, tension, pH, calcium, voltage and other variables. In some
sense one could consider the overall membrane as a sensitive
receptor for all of these variables if one is close to a transition.
9.1 Comparison with proteins
When comparing the lipid ion channels with protein channels
it has to be stated that
• typical conductances and lifetimes of protein and lipid
channels are of similar magnitude. The two classes of
events are seemingly indistinguishable.
If two events cannot be distinguished one has to reconsider
what the factual evidence actually is that allows to attribute
the current events to different processes. The findings on lipid
channels reported here imply that
1. the observation of quantized currents is no proof for the
action of ion channel proteins.
2. the action of a drug on a conduction event is no proof for
specific binding of the drug to a channel protein.
In logics, elementary rules of inference distinguish between
necessary and sufficient criteria. For channel proteins it must
be concluded that the finding of quantized currents alone is
neither necessary nor sufficient to infer the action of a receptor.
The quantized current could be though the lipid membrane,
and there is no particular reason why the a quantization should
indicate the action of a macromolecule. Similarly, the action of
drugs is not sufficient to infer specific binding. The drug could
have an effect without binding, as shown in section 6.3 for the
anesthetics. For example, the inference that the abolition of
channel events in biomembranes by tetrodotoxin (TTX) proves
the action of a sodium channel protein and the specific binding
of TTX to this channel is not admissible without additional
information, which is typically not provided in the literature.
The experiments reviewed in this paper do not give an in-
dication for that protein channels do not do exactly what has
been suggested in the protein ion channel literature. However,
the data provide considerable evidence that one can find the
same events in the absence of proteins and therefore one does
not necessarily require proteins as an explanation. While the
existence of lipid channels is easy to prove because a lipid
membrane can be measured in the absence of proteins, it is
presently impossible to measure protein channel conductances
in the absence of a membrane. Thus, it is much more difficult
Figure 27: Schematic drawing of proteins as perturbation
of the lipid membrane state. Top: Protein with long hy-
drophobic length embedded into a fluid lipid membrane.
Bottom: Protein with short hydrophobic length embedded
into a gel lipid membrane. Due to wetting of the protein
interface the lipids close to the protein are in a transition
state with high fluctuations and an increased likelihood
of lipid channel formation.
length and is surrounded by a gel membrane of larger thickness,
the protein interface will constitute a location for lipid channel
formation (Fig. 27, bottom). It is a general predictions from
such simulations that the tendency of proteins to aggregate and
their potential to trigger lipid ion channel formation are closely
coupled.
The role of the protein in such a scenario can be seen as dop-
ing the membrane and influence the physics of its surrounding,
much like in semiconductors where the doping agent gallium
can influence the physics of silicon. However, the proteins in
this situation do not conduct ions themselves. Rather is the
outer interface of the protein the site of permeation. All argu-
ments arguing in favor of a selectivity3 of a protein pore (chan-
nel protein) would also be true for the outer interface. Thus,
the protein in the above situation could be considered as an in-
verted pore (the pore interface is on the outside). The physics,
however, is that of the lipid membrane.
9 Discussion
We have reviewed here some of the literature indicating the
presence of quantized channel events in lipid membranes. It is
3According to Einstein [89] an all-or-nothing selectivity of a membrane (or
a membrane protein) for a particular substance violates the laws of thermody-
namics. Therefore, we rather refer to different permeation rates.
21
to actually prove that a protein ion channel is really active in an
experiment. From a patch-clamp experiment it is impossible to
conclude, which path a particular current event has taken. The
patch area is between 1-100 µm2 while the cross section of a
single channel protein, e.g. the potassium channel, is rather 25
nm2 = 25·10−6µm2. A typical aqueous pore has a diameter of
0.5 nm. Since no experiment can actually monitor the path of
the ions and therefore demonstrate the existence of a particular
pore inside a membrane protein, the argument is typically of
rather indirect nature:
• the protein increases the probability to find conduction
events, and conductivities are small in the absence of the
proteins
• the currents can be blocked by specific drugs as
tetrodotoxin (TTX) or tetraethylamonium (TEA)
• site specific mutation can alter or block the currents
One must state that it has been incorrectly assumed that the
lipid membrane is inert and does not display quantized con-
duction events. Second, proteins can act as perturbations of the
membrane and thus have the potential of altering the state of
the membrane. A direct proof for this effect is still missing, but
it is straight-forward to conclude this from the thermodynamics
described in this article. Therefore, a conductance that displays
a protein concentration dependence must be taken as a proof
for a correlation of the proteins with the conduction events but
not as a proof for the conductance through a pore in the protein
itself. Conduction events in lipid membranes depend also on
the concentration of drugs like anesthetics (as shown in section
6.3) even though anesthetic molecules do not form channels.
The dangers of confusing lipid with protein channels may be
evident in a publication by Cannon et al. [90]. These authors
found that calcium channels reconstituted in lipid mixtures are
not only most active at the phase boundaries of the lipid phase
diagram but also display the longest dwell times. While there
is no particular reason why a protein should behave this way,
this is exactly the behavior expected for lipid ion channels in
lipid membranes at phase boundaries (dwell times are long
in the transition, see section 7). The channel conductances in
this paper were the same as, e.g., in the lipid mixture shown in
Fig. 11. Further, MacKinnon and collaborators [91] showed
that the potassium channel conduction properties depend very
sensitively on the lipid environment proving a connection
between state of the lipid membrane and protein channels. The
authors assumed that the voltage sensors of these proteins are
lipid-controlled. The effect of anesthetics on lipid membrane
pores is very similar to that reported for the acetylcholine
receptor and many other channel proteins (reviewed in [51].
9.2 Nerve pulses
In a series of recent publications it has been shown that close
to melting transitions of lipid membranes one expects the
possibility of electromechanical pulse propagation (solitons) in
cylindrical membrane tubes [20, 92, 60, 93]. These solitons in
fact share many properties with the action potential in nerves
that cannot be explained by the Hodgkin-Huxley model, e.g.,
reversible heating of the nerve, and mechanical alterations like
thickness changes and length changes. Thus, it may be pos-
sible that the nerve pulse can be explained without involving
ion channel proteins. However, exactly under the conditions
under which such electromechanical solitons can propagate
one also expects the formation of lipid ion channels. As shown
in this article, everything that alters the the phase behavior will
alter the lipid channel appearance and simultaneously the free
energy necessary to induce a nervous pulse when following an
electromechanical approach to rationalize the phenomenology
of the action potential [20]. This is a beautiful correlation
purely based on the thermodynamics of the lipid membrane
that is well understood as shown here.
9.3 Conclusion
Taking the extreme relevance of the finding of quantized cur-
rents in artificial membrane where any effect of proteins can
be excluded it is quite surprising that this phenomenon has
not been studied more intensively. Not only do these currents
display similar conductances as proteins channels, but obvi-
ously similar fluctuation life times. Since many biomembranes
exist in a state close to melting transitions of their membranes,
the striking similarity between the lipid and protein events
suggests that these events are difficult or even impossible
to distinguish. Should in fact some of the reported protein
conductances be due to lipid pores, the theoretical description
of such channels would rather lie in the thermodynamics of
the membrane and its cooperative phase behavior than in the
geometry of individual proteins. A discussion of the lipid
channel phenomenon and its connection with protein channels
is overdue.
Even though there have been reports on lipid ion channels
for 30 years the investigation of lipid ion channels is still
in an early stage. Many of the thermodynamics couplings
described in this article have not been tested yet by experiment.
However, already now many of the implications of applying
the fluctuation-dissipation theorem to lipid channels found
experimental verification, e.g., the coupling between life times
and heat capacity, the effect of pH, anesthetic drugs and lateral
tension. Many further experiments are needed to get better
statistics of the membrane conductance and the influence of
various parameters. It would in particular be nice to demon-
strate the influence of proteins that cannot be channel proteins
(for instance peripheral proteins) on the quantized conductance
of lipid membranes. Such experiments are not difficult and
will most likely be done in the near future.
Acknowledgments:
I thank Matthias Fidorra, Katarzyna
Wodzinska, Andreas Blicher, Katrine R. Laub and Klaus B.
22
Pedersen who performed permeability measurements in my lab.
References
[1] A. L. Hodgkin, A. F. Huxley, A quantitative description of mem-
brane current and its application to conduction and excitation in
nerve, J. Physiol. 117 (1952) 500–544.
[2] A. Blicher, K. Wodzinska, M. Fidorra, M. Winterhalter, T. He-
imburg, The temperature dependence of lipid membrane perme-
ability, its quantized nature, and the influence of anesthetics, Bio-
phys. J. 96 (2009) 4581–4591.
[3] D. Papahadjopoulos, K. Jacobson, S. Nir, T. Isac, Phase tran-
sitions in phospholipid vesicles. fluorescence polarization and
permeability measurements concerning the effect of temperature
and cholesterol., Biochim. Biophys. Acta 311 (1973) 330–340.
[4] J. F. Nagle, H. L. Scott, Lateral compressibility of lipid mono-
and bilayers lateral compressibility of lipid mono- and bilayers.
theory of membrane permeability, Biochim. Biophys. Acta 513
(1978) 236–243.
[5] M. C. Sabra, K. J. rgensen, O. G. Mouritsen, Lindane sup-
presses the lipid-bilayer permeability in the main transition re-
gion, Biochim. Biophys. Acta 1282 (1996) 85–92.
[6] M. Jansen, A. Blume, A comparative study of diffusive and os-
motic water permeation across bilayers composed of phospho-
lipids with different head groups and fatty acyl chains, Biophys.
J. 68 (1995) 997–1008.
[7] M. Yafuso, S. J. Kennedy, A. R. Freeman, Spontaneous conduc-
tance changes, multilevel conductance states and negative differ-
ential resistance in oxidized cholesterol black lipid membranes,
J. Membr. Biol. 17 (1974) 201–212.
[8] V. F. Antonov, V. V. Petrov, A. A. Molnar, D. A. Predvoditelev,
A. S. Ivanov, The appearance of single-ion channels in unmodi-
fied lipid bilayer membranes at the phase transition temperature,
Nature 283 (1980) 585–586.
[9] G. Boheim, W. Hanke, H. Eibl, Lipid phase transition in planar
bilayer membrane and its effect on carrier- and pore-mediated
ion transport, Proc. Natl. Acad. Sci. USA 77 (1980) 3403–3407.
[10] K. Kaufmann, I. Silman, Proton-induced ion channels through
lipid bilayer membranes, Naturwissenschaften 70 (1983) 147–
149.
[11] K. Kaufmann, I. Silman, The induction by protons of ion chan-
nels through lipid bilayer membranes, Biophys. Chem. 18 (1983)
89–99.
[12] V. F. Antonov, E. V. Shevchenko, E. T. Kozhomkulov, A. A. Mol-
nar, E. Y. Smirnova, Capacitive and ionic currents in BLM from
phosphatidic acid in Ca2+-induced phase transition, Biochem.
Biophys. Research Comm. 133 (1985) 1098–1103.
[13] T. Heimburg, Thermal biophysics of membranes, Wiley VCH,
Berlin, Germany, 2007.
[14] M. Montal, P. Mueller, Formation of bimolecular membranes
from lipid monolayers and a study of their electrical properties.,
Proc. Natl. Acad. Sci. USA 69 (12) (1972) 3561–3566.
23
[15] E. Neher, B. Sakmann, Single-channel currents recorded from
membrane of denervated frog muscle fibres, Nature 260 (1976)
779–802.
[16] R. Coronado, R. Latorre, Phospholipid bilayers made from
monolayers on patch-clamp pipettes, Biophys. J. 43 (1983) 231–
236.
[17] W. Hanke, C. Methfessel, U. Wilmsen, G. Boheim, Ion chan-
nel reconstruction into lipid bilayer membranes on glass patch
pipettes, Bioelectrochem. Bioenerg. 12 (1984) 329–339.
[18] H. Hauser, M. C. Phillips, M. Stubbs, Ion permeability of phos-
pholipid bilayers, Nature 239 (1972) 342–344.
[19] P. Agre, M. Bonhivers, M. J. Borgnia, The aquaporins, blueprints
for cellular plumbing systems, J. Biol. Chem. 273 (1998) 14659–
14662.
[20] T. Heimburg, A. D. Jackson, On soliton propagation in biomem-
branes and nerves, Proc. Natl. Acad. Sci. USA 102 (2005) 9790–
9795.
[21] O. P. Hamill, A. Marty, E. Neher, B. Sakmann, F. J. Sigworth,
Improved patch-clamp techniques for high-resolution current
recording from cells and cell-free membrane patches, Pflugers
Arch. 391 (1981) 85–100.
[22] A. Bruggemann, M. George, M. Klau, M. Beckler, J. Steindl,
J. C. Behrends, N. Fertig, High quality ion channel analysis on a
chip with the npc technology, Assay Drug Dev. Technol. 1 (2003)
665–673.
[23] K. Yoshikawa, T. Fujimoto, T. Shimooka, H. Terada, N. Ku-
mazawa, T. Ishii, Electrical oscillation and fluctuations in phos-
pholipid membranes. phospholipids can form a channel without
protein, Biophys. Chem. 29 (1988) 293–299.
[24] M. Fidorra, Confocal microscopy, calorimetry and permeability
studies on giant lipid vesicles containing ceramides, Ph.D. thesis,
University of Southern Denmark (2007).
[25] T. M. Suchyna, V. S. Markin, F. Sachs, Biophysics and structure
of the patch and the gigaseal, Biophys. J. 97 (2009) 738–747.
[26] B. Hille, Ionic channels of excitable membranes, Cambridge
University Press, Cambridge, 1992.
[27] T. Heimburg, Die Physik von Nerven, Physik Journal 8 (3)
(2009) 33–39.
[28] D. J. Woodbury, Pure lipid vesicles can induce channel-like con-
ductances in planar bilayers, J. Membr. Biol. 109 (1989) 145–
150.
[29] K. Kaufmann, W. Hanke, A. Corcia,
Ion channels,
available at:
//membranes.nbi.dk/ Kaufmann/pdf/
Kaufmann book3 ed.pdf", Caruaru, library: Niedersachsische
Staats- und Universitatsbibliothek Gottingen, 37073 Gottingen,
Germany, call number 95 A 25498: Gottingen (1989).
"http:
[30] B. Wunderlich, C. Leirer, A. Idzko, U. F. Keyser, V. Myles,
T. Heimburg, M. Schneider, Phase state dependent current fluc-
tuations in pure lipid membranes, Biophys. J. 96 (2009) 4592–
4597.
[31] R. Benz, F. Beckers, U. Zimmermann, Reversible electrical
breakdown of lipid bilayer membranes: A charge-pulse relax-
ation study, J. Membr. Biol. 48 (1979) 181–204.
[32] I. G. Abidor, L. V. Chernomordik, S. I. Sukharev, Y. A. Chiz-
madzhev, The reversible electrical breakdown of bilayer lipid
membranes modified by uranyl ions, Bioelectrochem. Bioenerg.
9 (1982) 141–148.
[49] H. Ebel, P. Grabitz, T. Heimburg, Enthalpy and volume changes
in lipid membranes. i. the proportionality of heat and volume
changes in the lipid melting transition and its implication for the
elastic constants, J. Phys. Chem. B 105 (2001) 7353–7360.
[33] E. Neumann, S. Kakorin, K. Taensing, Fundamentals of electro-
porative delivery of drugs and genes, Bioelectrochem. Bioenerg.
48 (1999) 3–16.
[50] L. Cruzeiro-Hansson, O. G. Mouritsen, Passive ion permeability
of lipid membranes modelled via lipid-domain interfacial area.,
Biochim. Biophys. Acta 944 (1988) 63–72.
[34] V. F. Antonov, A. A. Anosov, V. P. Norik, E. Y. Smirnova,
Soft perforation of planar bilayer lipid membranes of dipalmi-
toylphosphatidylcholine at the temperature of the phase transi-
tion from the liquid crystalline to gel state, Eur. Biophys. J. 34
(2005) 155–162.
[35] P. A. Gurnev, D. Harries, V. A. Parsegian, S. M. Bezrukov,
The dynamic side of the hofmeister effect: a single-molecule
nanopore study of specific complex formation, Chem. Phys.
Chem. 10 (2009) 1445–1449.
[36] M. G. Cacace, E. M. Landau, J. J. Ramsden, The hofmeister
series: salt and solvent effects on interfacial phenomena, Quart.
Rev. Biophys. 30 (1997) 241–277.
[37] A. Finkelstein, Water movement through lipid bilayers, pores,
and plasma membranes. Theory and reality, Vol. 4 of Distin-
guished lecture series of the society of general physiologists, Wi-
ley, 1987.
[38] R. W. Glaser, S. L. Leikin, L. V. Chernomordik, V. F. Pas-
tushenko, A. I. Sokirko, Reversible breakdown of lipid bilayers:
Formation and evolution of pores, Biochim. Biophys. Acta 940
(1988) 275–287.
[39] E. Overton, Uber die osmotischen Eigenschaften der lebenden
Pflanzen und Thierzelle, Vierteljahresschr. Naturforsch. Ges.
Zurich 40 (1895) 159–201.
[40] R. Bockmann, R. de Groot, S. Kakorin, E. Neumann,
H. Grubmuller, Kinetics, statistics, and energetics of lipid mem-
brane electroporation studied by molecular dynamics simula-
tions, Biophys. J. 95 (2008) 1837–1850.
[41] D. G. Levitt, Kinetics of movement in narrow channels., Curr.
Top. Membr. Transp. 21 (1984) 181–198.
[42] M. Winterhalter, W. Helfrich, Effect of voltage on pores in mem-
branes, Phys. Rev. A 36 (1987) 5874–5876.
[43] A. Einstein, Theorie der Opaleszenz von homogenen
Flussigkeiten und Flussigkeitsgemischen in der Nahe des
kritischen Zustandes, Ann. Phys. (Leipzig) 33 (1910) 1275–
1298.
[44] L. Onsager, Reciprocal relations in irreversible processes. I.,
Phys. Rev. 37 (1931) 405–426.
[45] L. Onsager, Reciprocal relations in irreversible processes. II.,
Phys. Rev. 38 (1931) 2265–2279.
[46] D. Kondepudi, I. Prigogine, Modern thermodynamics, Wiley,
Chichester, 1998.
[47] R. F. Greene, H. B. Callen, On the formalism of thermodynamic
fluctuation theory, Phys. Rev. 83 (1951) 1231–1235.
[48] T. Heimburg, Mechanical aspects of membrane thermodynam-
ics. Estimation of the mechanical properties of lipid membranes
close to the chain melting transition from calorimetry, Biochim.
Biophys. Acta 1415 (1998) 147–162.
[51] K. Wodzinska, A. Blicher, T. Heimburg, The thermodynamics of
lipid ion channel formation in the absence and presence of anes-
thetics. blm experiments and simulations., Soft Matter 5 (2009)
3319–3330.
[52] E. Corvera, O. G. Mouritsen, M. A. Singer, M. Zuckermann,
The permeability and the effect of acyl-chain length for phos-
pholipid bilayers containing cholesterol: theory and experiment.,
Biochim. Biophys. Acta 1107 (1992) 261–270.
[53] T. Heimburg, A. D. Jackson, The thermodynamics of general
anesthesia, Biophys. J. 92 (2007) 3159–3165. doi:doi:10.
1529/biophysj.106.099754.
[54] H. Meyer, Zur Theorie der Alkoholnarkose. Erste Mittheilung.
Welche Eigenschaft der Anasthetica bedingt ihre narkotische
Wirkung?, Arch. Exp. Pathol. Pharmakol. 425 (1899) 109–118.
[55] C. E. Overton, Studien uber die Narkose, Verlag Gustav Fischer.,
1901. Jena, Germany. English translation: Studies of Narcosis,
Chapman and Hall, 1991, R. Lipnick, Ed.
[56] C. E. Overton, Studies of narcosis, Chapman and Hall, 1991,
english version of 'Studien der Narkose' from 1901.
[57] H. M. Seeger, M. L. Gudmundsson, T. Heimburg, How anes-
thetics, neurotransmitters, and antibiotics influence the relax-
ation processes in lipid membranes, J. Phys. Chem. B 111 (2007)
13858–13866.
[58] J. E. Rothman, J. Lenard, Membrane asymmetry, Science 195
(1977) 743–753.
[59] J. E. Rothman, E. P. Kennedy, Asymmetrical distribution of
phospholipids in the membrane of bacillus megaterium, J. Mol.
Biol. 110 (1977) 603–618.
[60] T. Heimburg, A. D. Jackson, On the action potential as a prop-
agating density pulse and the role of anesthetics, Biophys. Rev.
Lett. 2 (2007) 57–78.
[61] H. Trauble, M. Teubner, P. Woolley, H. Eibl, Electrostatic inter-
actions at charged lipid membranes. i. effects of ph and univalent
cations on membrane structure., Biophys. Chem. 4 (1976) 319–
342.
[62] V. F. Antonov, E. Y. Smirnova, E. V. Shevchenko, Electric field
increases the phase transition temperature in the bilayer mem-
brane of phosphatidic acid., Chem. Phys. Lipids 52 (1990) 251–
257.
[63] A. Jakli, J. Harden, C. Notz, C. Bailey, Piezoelectricity of
phospholipids: a possible mechanism for mechanoreception and
magnetoreception in biology, Liquid Cryst. 35 (2008) 395–400.
[64] K. Y. C. Lee, H. M. McConnell, Effect of electric-field gradients
on lipid monolayer membranes, Biophys. J. 68 (1995) 1740–
1751.
[65] H. Gogelein, H. Koepsell, Channels in planar bilayers made from
commercially available lipids, Pflugers Arch. 401 (1984) 433–
434.
24
[66] T. Baumgart, S. T. Hess, W. W. Webb, Imaging coexisting fluid
domains in biomembrane models coupling curvature and line
tension, Nature 425 (2003) 821–824.
[82] T. Heimburg, R. L. Biltonen, A Monte Carlo simulation study
of protein-induced heat capacity changes, Biophys. J. 70 (1996)
84–96.
[83] V. P. Ivanova, I. M. Makarov, T. E. Schaffer, T. Heimburg, Anal-
izing heat capacity profiles of peptide-containing membranes:
Cluster formation of gramicidin A, Biophys. J. 84 (2003) 2427–
2439.
[84] M. Fidorra, Untersuchung des Phasenverhaltens von Membranen
durch konfokale Mikroskopie und Kalorimetrie., Master's thesis,
University of Gottingen (2004).
[85] H. Seeger, M. Fidorra, T. Heimburg, Domain size and fluctua-
tions at domain interfaces in lipid mixtures., Macromol. Sym-
posia 219 (2005) 85–96.
[86] E. Freire, T. Markello, C. Rigell, P. W. Holloway, Calorimet-
ric and fluorescence characterization of interactions between cy-
tochrome b5 and phosphatidylcholine bilayers, Biochemistry 28
(1983) 5634–5643.
[87] M. R. Morrow, J. H. Davis, F. J. Sharom, M. P. Lamb, Studies
of the interaction of human erythroyte band 3 with membrane
lipids using deuterium nuclear magnetic resonance and differ-
ential scanning calorimetry, Biochim. Biophys. Acta 858 (1986)
13–20.
[88] O. G. Mouritsen, M. Bloom, Mattress model of lipid-protein in-
teractions in membranes., Biophys. J. 46 (1984) 141–153.
[89] A. Einstein, uber die thermodynamische Theorie der Poten-
tialdifferenz zwischen Metallen und vollstandig dissociirten
Losungen ihrer Salze und uber eine elektrische Methode zur Er-
forschung der Molekularkrafte, Ann. Phys. (Leipzig) 8 (1902)
798–814.
[90] B. Cannon, M. Hermansson, S. Gyorke, P. Somerharju, J. A.
Virtanen, Regulation of calcium channel activity by lipid domain
formation in planar lipid bilayers, Biophys. J. 85 (2003) 933–
942.
[91] D. Schmidt, Q.-X. Jiang, R. MacKinnon, Phospholipids and
the origin of cationic gating charges in voltage sensors., Nature
444 (7120) (2006) 775–779. doi:10.1038/nature05416.
[92] B. Lautrup, A. D. Jackson, T. Heimburg, The stability of solitons
in biomembranes and nerves, arXiv:physics/0510106.
[93] S. S. L. Andersen, A. D. Jackson, T. Heimburg, Towards a ther-
modynamic theory of nerve pulse propagation, Progr. Neurobiol.
88 (2009) 104–113.
[67] W. Burack, Q. Yuan, R. Biltonen, Role of lateral phase separation
in the modulation of phospholipase A2 activity, Biochemistry 32
(1993) 583–589.
[68] F. H. Johnson, E. A. Flagler, Hydrostatic pressure reversal of
narcosis in tadpoles., Science 112 (1950) 91–92.
[69] A. Punnia-Moorthy, Evaluation of ph changes in inflammation of
the subcutaneous air pouch linig in the rat, induced by carreenan,
dextran and staphylococcus aureus, J. Oral Pathol. 16 (1987) 36–
44.
[70] T. Y. Tsong, Kinetics of the crystalline-liquid crystalline phase
transition of dimyristoyl l-lecithin bilayers, Proc. Natl. Acad.
Sci. USA 71 (1974) 2684–2688.
[71] B. Gruenewald, A. Blume, F. Watanabe, Kinetic investigations
on the phase transition of phospholipid bilayers., Biochim. Bio-
phys. Acta 597 (1980) 41–52.
[72] K. Elamrani, A. Blume, Phase transition kinetics of phosphatidic
acid bilayers. a pressure-jump relaxation study, Biochemistry 22
(1983) 3305 – 3311.
M.
bilayers.
Dimyristoylphosphatidic
acid/cholesterol
and
kinetics of the phase transition as studied by the pressure jump
relaxation technique, Europ. Biophys. J. 13 (1986) 343–353.
Hillmann,
thermodynamic
properties
[73] A.
Blume,
[74] W. W. van Osdol, R. L. Biltonen, M. L. Johnson, Measuring
the kinetics of membrane phase transition, J. Bioener. Biophys.
Methods 20 (1989) 1–46.
[75] W. W. van Osdol, M. L. Johnson, Q. Ye, R. L. Biltonen, Re-
laxation dynamics of the gel to liquid crystalline transition of
phosphatidylcholine bilayers .effects of chainlength and vesicle
size, Biophys. J. 59 (1991) 775–785.
[76] P. Grabitz, V. P. Ivanova, T. Heimburg, Relaxation kinetics of
lipid membranes and its relation to the heat capacity, Biophys. J.
82 (2002) 299–309.
[77] J. Gallaher, K. Wodzinska, T. Heimburg, M. Bier, Ion-channel-
like behavior in lipid bilayer membranes at the melting transi-
tion, Submitted.
[78] V. Oliynyk, U. Kaatze, T. Heimburg, Pore formation of lytic pep-
tides in lipid membranes and their influence on the thermody-
namic properties of the pore environment., Biochim. Biophys.
Acta 1768 (2007) 236–245.
[79] O. G. Mouritsen, Computer simulation of cooperative phenom-
ena in lipid membranes, in: R. Brasseur (Ed.), Molecular de-
scription of biological membrane components by computer aided
conformational analysis, CRC Press, Boca Raton, 1990, pp. 3–
83.
[80] O. G. Mouritsen, M. M. Sperotto, J. Risbo, Z. Zhang, M. J. Zuck-
ermann, Computational approach to lipid-protein interactions in
membranes, in: H. Villar (Ed.), Advances in Computational Bi-
ology, JAI Press, Greenwich, Conneticut, 1995, pp. 15–64.
[81] O. G. Mouritsen, K. Jorgensen, Micro-, nano- and meso-scale
heterogeneity of lipid bilayers and its influence on macroscopic
membrane properties, Mol. Membr. Biol. 12 (1995) 15–20.
25
|
1206.4841 | 2 | 1206 | 2013-02-19T14:25:30 | In-phase and anti-phase synchronization in noisy Hodgkin-Huxley neurons | [
"physics.bio-ph",
"q-bio.NC"
] | We numerically investigate the influence of intrinsic channel noise on the dynamical response of delay-coupling in neuronal systems. The stochastic dynamics of the spiking is modeled within a stochastic modification of the standard Hodgkin-Huxley model wherein the delay-coupling accounts for the finite propagation time of an action potential along the neuronal axon. We quantify this delay-coupling of the Pyragas-type in terms of the difference between corresponding presynaptic and postsynaptic membrane potentials. For an elementary neuronal network consisting of two coupled neurons we detect characteristic stochastic synchronization patterns which exhibit multiple phase-flip bifurcations: The phase-flip bifurcations occur in form of alternate transitions from an in-phase spiking activity towards an anti-phase spiking activity. Interestingly, these phase-flips remain robust in strong channel noise and in turn cause a striking stabilization of the spiking frequency. | physics.bio-ph | physics |
In-phase and anti-phase synchronization in noisy
Hodgkin-Huxley neurons
Xue Ao, Peter Hanggi, Gerhard Schmid
Institut fur Physik, Universitatsstr. 1, 86159 Augsburg, Germany
Abstract
We numerically investigate the influence of intrinsic channel noise on the dy-
namical response of delay-coupling in neuronal systems. The stochastic dy-
namics of the spiking is modeled within a stochastic modification of the stan-
dard Hodgkin-Huxley model wherein the delay-coupling accounts for the finite
propagation time of an action potential along the neuronal axon. We quan-
tify this delay-coupling of the Pyragas-type in terms of the difference between
corresponding presynaptic and postsynaptic membrane potentials. For an el-
ementary neuronal network consisting of two coupled neurons we detect char-
acteristic stochastic synchronization patterns which exhibit multiple phase-flip
bifurcations: The phase-flip bifurcations occur in form of alternate transitions
from an in-phase spiking activity towards an anti-phase spiking activity.
In-
terestingly, these phase-flips remain robust in strong channel noise and in turn
cause a striking stabilization of the spiking frequency.
Keywords:
delayed coupling
synchronization, channel noise, stochastic Hodgkin-Huxley,
1. Introduction
Time-delayed feedback presents a common mechanism which is found in
many biological systems including neuronal systems. Signal transmission time
delays in neuronal systems either result from (i) chemical processes in the neu-
ronal synapses where neurotransmitters are released and diffusively overcome
the synaptic cleft and/or (ii) from the finite propagation speed of electrical exci-
tations along the neuronal axon. Time delays stemming from chemical synapses
are of the order of a few milliseconds, while the axonal conduction delays in both,
delay-coupled neurons and autaptic feedback loops, reach values up to tens of
milliseconds [1 -- 4].
As the time scale of the delayed coupling and of the neuronal dynamics
become comparable, the delay-coupling gives raise to peculiar synchronization
Email address: [email protected] (Gerhard Schmid)
Preprint submitted to Elsevier
November 8, 2018
In particular, phase synchronization phenomena in neuronal
phenomena [5].
systems are commonly thought to be the basis for many biological relevant pro-
cesses occurring in the brain [6, 7]. Synchrony of neurons from small brain
regions up to large-scale networks of different cortices comes along with trans-
mission time delays. Theoretical and computational studies on neuronal net-
works with delay-coupling recently highlighted the occurrence of so-called phase-
flip bifurcations [8 -- 10]. The ensemble activity of the coupled neurons change
abruptly from in-phase to anti-phase oscillations or vice versa.
With this work we research this objective by considering the influence of
internal noise. It is an established fact that noise leads to various prominent
effects in neuronal dynamics [11]. Some typical examples that come to mind are
stochastic resonance features [12 -- 15], and noise-assisted synchronization [5, 16 --
18]. Within our work the intrinsic noise is due to the stochastic gating of the
ion channels, i.e. the so-called channel noise which is inherently coupled to the
electrical properties of the axonal cell membrane [19 -- 21]. Interestingly, it has
been shown that intrinsic channel noise does not only lead to the generation
of spontaneous action potentials [22], but as well affects the neuronal dynam-
ics at different levels, namely: (i) it can boost the signal quality [14, 15], (ii)
enhance the signal transmission reliability [23], (iii) cause frequency- and phase-
synchronization features [24 -- 28] and (iv) may result in a frequency stabilization
[29], to name but a few.
The present work is organized as follows: In Sec. 2 we introduce the biophys-
ical model. We review the standard Hodgkin-Huxley model and its generaliza-
tions with respect to intrinsic channel noise and a delay-coupling. Numerical
methods for simulation are introduced after that. In Sec. 3, the dynamics of a
network of two delay-coupled Hodgkin-Huxley neurons is explored both, in the
deterministic limit and under consideration of channel noise. As a comparison,
we retrospect on the previous work on a single neuron subjected to a delayed
feedback loop resulting from autapse in Sec. 4. Our conclusions are given in
Sec. 5.
2. Biophysical model setup
We consider a minimal building block of a neuronal network composed of two
coupled neurons. As an archetype model for nerve excitation of the individual
neuron, we utilize a stochastic generalization of the common Hodgkin-Huxley
model. The stochastic generalization accounts for intrinsic membranal con-
ductance fluctuations, i.e. channel noise, being caused by random ion channel
gating. Moreover, we account for a delay in the coupling which accounts for a
finite propagation time of the action potential along the axon.
2.1. Hodgkin-Huxley-type modelling of two delay-coupled neurons
According to Hodgkin and Huxley, the dynamics of the membrane potential
Vi with i = 1, 2 of two coupled neuronal cells is given by [30]
dt Vi + GK(n) (Vi − VK)
C d
2
+GNa(m, h) (Vi − VNa) + GL (Vi − VL) = Ii(t) .
(1)
Here, Vi denotes the membrane potential of the i-th cell. The stimulus Ii(t)
acting on the i-th neuron reads:
Ii(t) = Ii,ext(t) + I τ
i,j(t) ,
i, j = 1, 2 , i (cid:54)= j ,
(2)
where the bi-directional delay-coupling of Pyragas-type [31] between the two
neurons is assumed to be linear in the difference of the membrane potentials of
a primary, i-th neuron at time t and a secondary, j-th neuron at an earlier time,
t − τ . The coupling thus reads:
i,j(t) = κ [Vj(t − τ ) − Vi(t)] ,
I τ
(3)
where κ corresponds to the coupling strength and τ denotes the finite delay
time. The coupling defined in Eq. (3) is of "electrotonic" type, i.e. we consider
an idealized situation wherein the coupling is proportional to the difference of
presynaptic and postsynaptic membrane potentials. This kind of coupling then
corresponds to so-called gap-junctions which allow the bi-directional transport
of ions and small molecules from one neuronal cell into another. Unlike the
conductance of chemical synapses, the conductance of gap-junctions is indepen-
dent of the presynaptic and postsynaptic membrane potential and can therefore
be modelled by the constant coupling parameter κ. Possible chemical mecha-
nisms occurring at the synaptic cleft are assumed to be instantaneous as the
time scale for signal propagation along the neurons axon is much larger than
the corresponding one for the transport process in the synaptic cleft. Note, that
the delayed stimulus in Eq. (3) results in an excitatory coupling mechanism in
which the spiking of neuron i at an earlier time t − τ time favors the initiation
of a action potential of the other cell at time t.
In addition to the delayed, bilinear coupling current we apply a constant
current stimulus Ii,ext on the neurons, mimicking the common stimulus of
the neuronal environment on the so considered two-neuron network.
In ab-
sence of the bi-directional coupling the dynamics of each neuron exhibits a
bifurcation scenario exhibiting a subcritical Hopf bifurcation. As a conse-
quence, the membranal dynamics displays (i) a stable fix-point, i.e.
the so-
called rest state for Ii,ext < I1 ≈ 6.26µA/cm2, (ii) a stable spiking solution for
Ii,ext > I2 ≈ 9.763µA/cm2 and (iii) a bistable regime for which the stable rest
state and a stable oscillatory spiking solution coexist, i.e. for I1 < Ii,ext < I2
[32 -- 36]. In particular, for Ii,ext = 0 the membrane potential is Vrest = −65.0mV.
Throughout this work the membrane potentials are measured in units of mV
and time in units of ms. For a squid giant axon, the parameters in Eq. (1) read
VNa = 50 mV, VK = −77 mV, VL = −54.4 mV, and C = 1 µF/cm2. Further-
more, the leakage conductance is assumed to be constant, GL = 0.3 mS/cm2.
On the contrary, the sodium and potassium conductances are controlled by the
voltage-dependent gating dynamics of single ion channels and are proportional
to their respective numbers. In the Hodgkin-Huxley model [30], the opening of
the potassium ion channel is governed by four identical activation gates, being
3
characterized by the opening probability n. The channel is open when all four
gates are open. In the case of sodium channel, the dynamics is governed by a set
of three independent and identical fast activation gates (m) and an additional
slow, so-termed inactivation gate (h). The independence of the gates implies
that the probability PK,Na of the occurrence of the conducting conformation is
PK = n4 for a potassium channel and PNa = m3 h for a sodium channel, re-
spectively. In a mean field description, the macroscopic potassium and sodium
conductances then read:
GK(n) = gmax
K
n4, GNa(m, h) = gmax
Na m3h ,
(4)
K = 36 mS/cm2 and gmax
where gmax
Na = 120 mS/cm2 denote the maximal conduc-
tances (when all channels are open). The two-state, opening -- closing dynamics
of the gates is governed by the voltage dependent opening and closing rates
αx(V ) and βx(V ) (x = m, h, n), i.e., [30]
0.01 (V + 55)
1 − exp [− (V + 55) /10]
,
αn(V ) =
βn(V ) = 0.125 exp [− (V + 65) /80] ,
0.1 (V + 40)
αm(V ) =
1 − exp [− (V + 40) /10]
,
βm(V ) = 4 exp [− (V + 65) /18] ,
αh(V ) = 0.07 exp [− (V + 65) /20] ,
1 + exp [− (V + 35) /10]
.
βh(V ) =
1
(5)
(6)
(7)
(8)
(9)
(10)
Hence, the dynamics of the opening probabilities for the gates read:
x = αx(V ) (1 − x) − βx(V ) x,
x = m, h, n .
(11)
The voltage equation (1), Eq. (4) and the rate equations of the gating dynam-
ics Eqs. (5)- (11) then constitute the original, strictly deterministic Hodgkin --
Huxley model for spiking activity of the squid giant axon.
2.2. Modelling Channel Noise
In this study, however, each channel defines a bistable stochastic element
which fluctuates between its closed and open states. As a consequence, the
number of open channels undergoes a birth-death stochastic process. Applying
a diffusion approximation to this discrete dynamics, the resulting Fokker-Planck
equation can be obtained from a Kramers-Moyal expansion [37, 38]. The cor-
responding Langevin dynamics, interpreted here in the stochastic Ito calculus
[39], reads:
x = αx(V ) (1 − x) − βx(V ) x + ξx(t) ,
d
dt
x = n, m, h .
(12)
It is driven by independent Gaussian white noise sources ξx(t) of vanishing
mean which account for the fluctuations of the number of open gates. The
4
(multiplicative) noise strengths depend on both, the membrane voltage and the
gating variables. Explicitly, these noise correlations assume the following form
for a neuron consisting of NNa sodium and NK potassium ion channels:
(cid:104)ξm(t)ξm(t(cid:48))(cid:105) =
(cid:104)ξh(t)ξh(t(cid:48))(cid:105) =
(cid:104)ξn(t)ξn(t(cid:48))(cid:105) =
NN a
(1 − m)αm + mβm
(1 − h)αh + hβh
(1 − n)αn + nβn
NN a
NK
δ(t − t(cid:48)),
δ(t − t(cid:48)),
δ(t − t(cid:48)).
(13)
(14)
(15)
The fluctuations of the number of open ion channels result in conductances
fluctuations of the cell membrane eventually leading to spontaneous action po-
tentials. These spontaneous spiking events occur even for sub-threshold, con-
stant external current stimuli, i.e. for Ii,ext < I1. If the times of spike occur-
rences are given by tn with n = 0, 1, 2, ..., N , where N is the number of observed
spikes, the interspike interval between two succeeding spike is Tn = tn − tn−1
(n = 1, .., N ). For the case of a single Hodgkin-Huxley neuron the distribution
of these interspike intervals exhibits a peak-like structure with the peak located
around the intrinsic time Tintrinsic, which is determined by the limit cycle of the
deterministic dynamics [22].
The strength of the channel noise scales inversely with the number of the
ion channels. Consequently, the threshold for excitation can be reached more
easily with increasing the noise strength (i.e. smaller system size). In order to
characterize the spontaneous spiking, we introduce the mean interspike interval
(cid:104)T(cid:105) := lim
N→∞
1
N
N(cid:88)
n=1
Tn .
(16)
With increasing number of ion channels, i.e. decreasing channel noise level, the
mean interspike interval increases exponentially for a vanishing current stimulus
Iext = 0 [22].
In presence of a finite positive constant current stimulus Iext, the mean inter-
spike interval is always smaller than that for the undriven case [40]. Moreover,
for supra-threshold driving, i.e. Iext > I2, noise-induced skipping of spikes is
observed. Accordingly, the channel noise does not only favor the generation of
spikes, but can as well suppress deterministic spiking. For intermediate constant
current driving, i.e. I1 ≤ Iext ≤ I2, for which the Hodgkin-Huxley model ex-
hibits a bi-stability between a spiking and a non-spiking solution, channel noise
results in transitions between these two states.
2.3. Numerical Methods
Our numerical results are obtained via the numerical integration of the
stochastic dynamical system given by Eqs. (1)-(12). Particularly, we apply the
stochastic Euler-algorithm in order to integrate the underlying stochastic dy-
namics [41]. An integration step ∆t = 0.01 ms has been used in the simulations;
5
for the generation of the Gaussian distributed random numbers, the Box-Muller
algorithm [42] has been employed. The occurrence of a spiking event in the volt-
age signal Vi(t) is obtained by upward-crossing of a detection threshold value
V0 = 0. It turns out that this threshold can be varied over a wide range with
no influence on the resulting spike train dynamics.
To ensure that throughout all times the non-negative gating variables take
on values solely between 0 (all gates are closed) and 1 (all gates are open), we
implemented numerically reflecting boundaries at 0 and 1. Throughout this
work we assume a constant ratio of the numbers of potassium and sodium chan-
nels which results from constant ion channel densities. For this work we have
assumed channel densities of 20 potassium channels and 60 sodium channels per
µm2.
In performing the numerics we initially prepare each neuron in the rest state
voltage value Vrest. By applying a short current pulse on one of the two neurons,
we initialize an action potential in this neuron which later on can be echoed by
the delay-coupling.
3. Synchronization
In order to investigate the temporal correlation between the spiking statistics
of the two neurons, we apply the linearly interpolated, instantaneous time-
dependent phase Φi(t) of a stochastic spiking process of neuron i (i = 1, 2)
[17, 18]; i.e., for t ∈ [ti,n, ti,n+1]
Φi(t) = 2π n + 2π
t − ti,n
ti,n+1 − ti,n
, n = 0, 1, 2, ...Ni − 1 ,
(17)
where ti,n denotes the n-th spiking of neuron i and Ni is the total number of
spike events in the dynamics of neuron i. Note, that each spike occurrence
contributes to the overall phase with 2π. Between two succeeding spikes the
phase is obtained by linear interpolation.
Note, that in accordance with this definition of the phase, the angular spiking
frequency ωi is given by:
ωi = lim
t→∞
Φi(t)
t
.
(18)
This frequency ωi and the mean spiking rate being the inverse of the mean
interspike interval (cid:104)T(cid:105)i in Eq. (16) are equivalent up to a constant factor; i.e.
we obtain,
.
=
ωi
2π
(cid:104)T(cid:105)i
.
(19)
Due to this equivalence we present our simulation results solely in terms of the
mean interspike interval (cid:104)T(cid:105).
In the pursuit for a measure for the synchronization of spike occurrences in
the two coupled neurons, we consider the phase difference between the two spike
trains, namely:
∆Φ(t) =
Φ1(t) − Φ2(t)
mod 2π .
(20)
(cid:16)
(cid:17)
6
(color online) The interspike interval (cid:104)T(cid:105) for the spiking of one of the two delay-
Figure 1:
coupled standard Hodgkin-Huxley neurons with no external constant current stimulus, i.e.
Iext = 0, as function of the delay time τ for coupling strength κ = 0.2 mS/cm2 in panel (a)
and as function of both parameters of the delay-coupling, i.e. the coupling strength κ and the
delay time τ , in panel (b). The color encoding for the phase diagram in panel (b) is shown in
the color bar: the darker the color the shorter is the interspike interval (cid:104)T(cid:105). For white regions,
the mean interspike interval (cid:104)T(cid:105) tends to infinity and no repitive firing is observed.
It turns out that the phase difference ∆Φ(t) numerically tends, after transient
effects have faded away, for t → ∞ to a value that depends on various parameters
such as the coupling strength, the delay time, the total integration time and to
some extent (due to inherent irregular, chaotic behavior) even on the time step
used in our integration step. Accordingly, we observe in the long time limit that
ω1 = ω2 and (cid:104)T(cid:105)1 = (cid:104)T(cid:105)2 =: (cid:104)T(cid:105).
3.1. Deterministic limit
We consider first the deterministic limit by letting NNa → ∞ and NK → ∞.
In doing so, the occurrence of repetitive firing was systematically analyzed by
varying the two coupling parameters, i.e. the coupling strength κ and the delay
time τ .
For Iext = 0, i.e. for a subthreshold constant current driving, the resulting
mean interspike interval (cid:104)T(cid:105) is depicted in Fig. 1(a). For the coupling pa-
rameters taken within the white region of Fig. 1(b), the system relaxes to the
non-spiking rest state. However, in the regime of the spiking dynamics a spike
in any of the two neurons generates a subsequent spike in the other neuron.
Consequently, repetitive, but alternating, firing can be observed for both neu-
rons. The mean interspike interval (cid:104)T(cid:105) increases linearly with increasing delay
time τ , cf. Fig. 1(a). In particular,
(21)
where Tact ≈ 2ms is the activiation time which is the time between the time
the stimulus of delayed coupling sets in and the occurrence of the stimulated
(cid:104)T(cid:105) ≈ 2(Tact + τ ) ,
7
Figure 2:
(color online) The phase difference ∆Φ, cf. Eq. (20), for two delay-coupled standard
Hodgkin-Huxley neurons for Iext = 0, cf. Eqs. (1)-(11), is depicted as function of the delay
time τ in panel (a). In panel (b) the phase difference ∆Φ is depicted within a phase diagram
upon varying the two coupling parameters κ and τ . In the white region, there is no repetitive
firing, cf. Fig. 1. In the red region the two neurons fire alternatingly, resulting in a phase
difference of ∆Φ (cid:39) π.
spiking. The factor '2' in Eq. (21) is due to the alternating spiking of the two
neurons. Note, that the time between succeeding spiking events of the network
is the delay time plus the activation time.
In Fig. 2 the steady-state phase difference ∆Φ is depicted as function of the
coupling strength κ and the delay time τ . For large delay times τ the neuronal
dynamics of an individual neuron possesses after each spiking event sufficient
time to relax back to its rest state before the delayed stimulus caused by the
spiking event of the other neuron sets in. This results in an alternating firing
dynamics of the two neurons. The spiking of the two neurons therefore exhibits a
constant phase shift of π, i.e. the spiking event of one neuron is almost perfectly
located between two succeeding spiking events in the other neuron. However, for
delay times that are of the order, or are even smaller than the refractory time,
an irregular behavior of the Hodgkin-Huxley dynamics emerges, as it is to be
expected in presence of finite delay. This in turn is reflected in the numerically
evaluated phase difference ∆Φ by a noisy behavior that succeedingly smoothes
for increasing large delay times, yielding the afore mentioned asymptotic phase
shift of π. This result is corroborated with our numerics, as depicted with Fig. 2.
3.2. Influence of channel noise
In Fig. 3 we depict the dependence of the mean interspike interval (cid:104)T(cid:105) on
the coupling parameter κ and delay time τ for three different levels of channel
noise, i.e. different sets of ion channel numbers NNa and NK. With increasing
noise level, i.e. decreasing number of ion channels, the sharp transition between
the parameter regime of repetitive spiking and non-repetitive spiking is smeared
out.
8
(color online) The interspike interval (cid:104)T(cid:105) for the spiking of one of the two delay-
Figure 3:
coupled standard Hodgkin-Huxley neurons with no external constant current stimulus, i.e.
Iext = 0, as function of the coupling parameters κ and the delay time τ . The mean interspike
interval (cid:104)T(cid:105) is depicted for three different strengths of the channel noise: (a) strong intrinsic
channel noise with NNa = 360 , NK = 120, (b) moderate intrinsic channel noise with NNa =
3600 , NK = 1200 and (c) weak intrinsic channel noise with NNa = 36000 , NK = 12000.
(color online) The steady-state phase difference ∆Φ for two delay-coupled stochastic
Figure 4:
Hodgkin-Huxley neurons for Iext = 0 is depicted as function of the delay coupling parameters
κ and τ for three different noise strengths: (a) strong intrinsic channel noise with NNa =
360 , NK = 120, (b) moderate intrinsic channel noise with NNa = 3600 , NK = 1200 and
(c) weak intrinsic channel noise with NNa = 36000 , NK = 12000. The stripes in panel (a)
indicate noise-induced phase-flips, see text.
9
Figure 5: (color online) In panel (a) the mean interspike interval (cid:104)T(cid:105) is depicted as function
of the coupling time τ for constant coupling strength κ, Iext = 0 and channel noise level
corresponding to numbers of ion channel: NNa = 360 and NK = 120. For the same parameters,
the phase difference ∆Φ as function of the delay time τ is depicted in (b).
In addition, for considerable strong channel noise, distinct synchronization
patterns emerge, indicating a frequency locking similar to the one observed for
the autaptic case discussed in Sec. 4 below. In order to analyze the observed
synchronization patterns in greater detail, we depict the mean interspike interval
(cid:104)T(cid:105) as function of the delay time τ for a fixed coupling strength κ = 0.7 mS/cm2
in Fig. 5(a). We find that the mean interspike interval varies with the delay time
in an almost piecewise linear manner, displaying sharp, triangle-like textures.
The distinct peak locations of the mean interspike interval (cid:104)T(cid:105) can be ex-
plained by the number of spikes that match in accordance with the intrinsic
time Tintrinsic a full propagation time length, given by twice the delay time. The
mean interspike interval (cid:104)T(cid:105) henceforth is proportional to the ratio of twice the
delay time and the number of spikes fitting into this very delay time interval, cf.
Eq. (21). Accordingly, channel noise results in a stabilization of the interspike
interval to an interval around the intrinsic time scale of the Hodgkin-Huxley
oscillator.
3.3. Phase-flip bifurcation
Note, that this phenomenon complies with similar pattern appearing in the
diagram for the phase difference, cf. Fig. 4. Furthermore, we depict the phase
difference ∆Φ in Fig. 5(b). These sharp transitions are accompanied by pro-
nounced noise-induced phase-flips. At the corresponding phase-flip values at 0
and π the spiking of the basic network changes from an in-phase towards an
anti-phase spiking: For ∆Φ ≈ 0, both neurons spike simultaneously, implying
synchronous firing, cf. Fig. 6.
10
Figure 6: (color online) The simulated time-course of the membrane potentials for the two
delayed coupled Hodgkin-Huxley neurons is depicted: (a) for the parameters κ = 0.7mS/cm2
and τ = 8ms,the dynamics shows alternating, i.e. antiphase, spiking, (b) for κ = 0.7mS/cm2
and τ = 15ms synchronous, inphase firing of the two neurons is observed.
4. Retrospect: Frequency locking by an autaptic feedback loop
The observed frequency stabilization in the dynamics of a network of two de-
layed coupled neurons shares it's origin with the frequency locking phenomenon
in noisy neurons with an autaptic feedback-loop. When neuronal dendrites es-
tablish an autapse, i.e. a connection to the neuron's own axon, a delayed feed-
back loop is induced to the neuron's dynamics. Such autosynapses, described
originally by Van der Loos and Glaser in 1972 [3] are a common phenomenon
found in about 80% of all analyzed pyramidal cells in the cerebral neocortex of
human brain [4].
Auto-synapses establish a time-delayed feedback mechanism on the cellular
level [3]. From a mathematical point of view, autaptic connections introduce
new time scales into the single neuron dynamics which gives raise to peculiar
frequency looking behavior [29]. Our modelling, cf. Eqs. (1)-(12), captures the
stochastic autapse phenomena for i = j and only one neuron.
In the limit of vanishing channel noise by letting NNa → ∞ and NK → ∞,
the spiking period is given by the delay time τ plus the activation time Tact,
being the time needed for creating the next spike event after the delayed stimulus
did set in, cf. Fig. 7(a). Note, that in presence of the autaptic delay, the fixed-
point solution of the unperturbed Hodgkin-Huxley dynamics remains stable and
the delayed stimulus is excitatory.
In presence of finite channel noise, however, there are two competitive mech-
anisms at work: (i) there are spiking repetitions due to the delay-coupling and
(ii) noise-induced generation of spikes or (iii) noise-induced skipping of spikes.
The interplay between these mechanisms becomes evident when the distribution
of the interspike intervals is considered. In particular, the interspike interval his-
togram (ISIH) exhibits a bimodal structure, exhibiting two peaks, see in Fig. 6
11
(color online) The interspike interval (cid:104)T(cid:105) for the standard Hodgkin-Huxley neuron
Figure 7:
with an autaptic feedback loop, cf. Eqs. (1)-(11). In panel (a) the dependence of the interspike
interval (cid:104)T(cid:105) on the delay time τ is depicted for different coupling strengths κ. In the regime
of repetitive firing the mean interspike interval (cid:104)T(cid:105) grows linearly with the delay time τ . In
the phase diagram shown in panel (b) the dependence of the interspike interval (cid:104)T(cid:105) on the
coupling strength κ and the delay time τ is depicted. The color bar next to panel (b) gives
the color encoding for the values of the interspike intervals. The white region corresponds
to the situation for which the externally initialized spike is not echoing itself and, formally,
(cid:104)T(cid:105) → ∞.
(color online) The mean interspike interval (cid:104)T(cid:105) for the stochastic Hodgkin-Huxley
Figure 8:
neuron with an autaptic feedback loop, cf. Eqs. (1)-(15). The channel noise level corresponds
to NNa = 360 and NK = 120. Similarly to Fig. 7, (cid:104)T(cid:105) is shown as function of the delay time
τ in panel (a), and the corresponding phase diagram is depicted in panel (b). The color bar
gives the color encoding of the mean interspike intervals (cid:104)T(cid:105).
12
in Ref. [29]. Upon specific values of the noise strength determined by the num-
ber of ion channels and the coupling parameters κ and τ the bimodal structure
shows distinct differences: Due to first mechanism the delay coupling leads to
a significant peak around the delay time τ . Via the noise-induced mechanisms,
the channel noise results in a broad peak around the intrinsic time scale Tintrinsic.
However, for considerable strong coupling strengths the bimodal structure
collapses to an unimodal one and a frequency-locking phenomena takes place.
The mean interspike interval (cid:104)T(cid:105) becomes bounded by a finite range of values
around the intrinsic time scale, but still shows a striking dependence on the
delay time τ , cf. Fig 8. In particular, the mean interspike interval (cid:104)T(cid:105) varies
with the delay time τ in an almost piecewise linear manner, displaying sharp
triangle-like textures, cf. Fig. 8(a) for κ = 0.7 mS/cm2 and τ > 10 ms.
5. Conclusions
With this work we have investigated the effects of intrinsic channel noise
on the spiking dynamics of an elementary neuronal network consisting of two
stochastic Hodgkin-Huxley neurons.
In doing so, we invoked some idealistic
simplifications such as the use of bilinear coupling with identical time delays
and equal coupling strengths. The finite transmission time of an action potential
traveling along the neuronal axon to the dendrites is the cause for the delay-
coupling to the other neuron. Physically this transmission time derives from
the finite length to the connecting dendrites and the finite propagation speed.
A Pyragas-like delayed feedback mechanism has been employed to model
the delayed coupling. The two basic parameters for the delay-coupling are the
coupling strength κ and the delay time τ . In terms of these two parameters the
delayed feedback mechanism results in a periodic, repetitive firing events of the
neurons.
Apart from this repetitive firing, the delay-coupling introduces intriguing
time scales. Particularly, we detect a noise-induced locking of the spiking rate
to the intrinsic frequency of the system. Consequently, the delayed feedback
mechanism serves as a control option for adjusting the interspike intervals; this
feature may be of importance for memory storage [43] and stimulus locked short-
term dynamics in neuronal systems [44]. One may therefore speculate whether
ubiquitous intrinsic channel noise in combination with the autapse phenomenon
is constructively harvested by nature for efficient frequency filtering.
Moreover, the emergence of a correlation between the firing dynamics of
the two delayed coupled neurons has been studied in terms of stochastic phase
synchronization. The dynamics of two coupled Hodgkin-Huxley neurons ex-
hibits noise-induced phase-flip bifurcations. At these phase-flip bifurcations the
phase difference changes abruptly and the spiking of the neuron switches from
an in-phase spiking to an anti-phase spiking, or vice versa. These phase-flips
are the direct result of the influence of channel noise. The observed phase-flips
thus may possibly assist the fact of a coexistence of various frequency rhythms
and oscillation patterns in different parts of extended neuronal networks, as for
example it is the case for the brain.
13
Acknowledgments
This work was supported by the Volkswagen foundation project I/83902 (PH,
GS) and by the German excellence cluster "Nanosystems Initiative Munich II"
(NIM II) via the seed funding project (XA, GS, PH).
References
References
[1] C. Masoller, M. C. Torrent, and J. Garc´ıa-Ojalvo, Phys. Rev. E 78, 041907
(2008).
[2] E. R. Kandel, J. H. Schwartz, and T. M. Jessell, Principles of Neural
Science, 3rd ed. (Elsevier, New York, 1991).
[3] H. Van Der Loos and E. M. Glaser, Brain Res. 48, 355 (1972).
[4] J. Lubke, H. Markram, M. Frotscher, and B. Sackmann, J. Neurosci. 16,
3209 (1996).
[5] A. Pikovsky, M. Rosenblum, and J. Kurths, Synchronization: A Universal
Concept in Nonlinear Sciences (Cambridge University Press, Cambridge,
England, 2003).
[6] F. Varela, J. P. Lachaux, E. Rodriguez,
and J. Martinerie, Nat. Rev.
Neurosci. 2, 229 (2001).
[7] J. P. Lachaux, E. Rodriguez, J. Martinerie, and F. Varela, Hum. Brain
Mapp. 8, 194 (1999).
[8] A. Prasad, J. Kurths, S. K. Dana, and R. Ramaswamy, Phys. Rev. E 74,
035204 (2006).
[9] A. Prasad, S. K. Dana, R. Karnatak, J. Kurths, B. Blasius, and R. Ra-
maswamy, Chaos 18, 023111 (2008).
[10] B. M. Adhikari, A. Prasad, and M. Dhamala, Chaos 21, 023116 (2011).
[11] B. Lindner, J. Garc´ıa-Ojalvo, A. Neimann, and L. Schimansky-Geier, Phys.
Rep. 392, 321 (2004).
[12] L. Gammaitoni, P. Hanggi, P. Jung, and F. Marchesoni, Rev. Mod. Phys.
70, 223 (1998).
[13] P. Hanggi, ChemPhysChem 3, 285 (2002).
[14] G. Schmid, I. Goychuk, and P. Hanggi, Europhys. Lett. 56, 22 (2001).
[15] P. Jung and J. W. Shuai, Europhys. Lett. 56, 29 (2001).
14
[16] A. M. Lacasta, F. Sagu´es, and J. M. Sancho, Phys. Rev. E 66, 045105(R)
(2002).
[17] L. Callenbach, P. Hanggi, S. J. Linz, J. A. Freund, and L. Schimansky-
Geier, Phys. Rev. E 65, 051110 (2002).
[18] J. A. Freund, L. Schimansky-Geier, and P. Hanggi, Chaos 13, 225 (2003).
[19] J. A. White, J. T. Rubinstein, and A. R. Kay, Trends Neurosci. 23, 131
(2000).
[20] J. H. Goldwin and E. Shea-Brown, PLOS Comput. Biol. 7, e1002247 (2011).
[21] G. Wainrib, M. Thieullen, and K. Pakdaman, J. Comput. Neursci. 32, 327
(2012).
[22] C. C. Chow and J. A. White, Biophys. J. 71, 3013 (1996).
[23] A. Ochab-Marcinek, G. Schmid, I. Goychuk, and P. Hanggi, Phys. Rev. E
79, 011904 (2009).
[24] G. Schmid, I. Goychuk, and P. Hanggi, Physica A 325, 165 (2002).
[25] J. M. Casado, Phys. Lett. A 310, 400 (2003).
[26] J. M. Casado and J. P. Baltan´as, Phys. Rev. E 68, 061918 (2003).
[27] B. Hauschildt, N. B. Janson, A. Balanov, and E. Scholl, Phys. Rev. E 74,
051906 (2006).
[28] L. Yu, Y. Chen, and P. Zhang, Eur. Phys. J. B 59, 249 (2007).
[29] Y. Li, G. Schmid, P. Hanggi, and L. Schimansky-Geier, Phys. Rev. E 82,
061907 (2010).
[30] A. L. Hodgkin and A. F. Huxley, J. Physiol. 117, 500 (1952).
[31] K. Pyragas, Phys. Lett. A 170, 421 (1992).
[32] P. Nelson, Biological Physics (W. H. Freeman and Company, New York,
2005).
[33] B. Hassard, J. Theor. Biol. 71, 401 (1978).
[34] J. Rinzel and R. N. Miller, Math. Biosci. 49, 27 (1980).
[35] G. Schmid, I. Goychuk, and P. Hanggi, Lect. Notes Phys. 625, 195 (2003).
[36] E. M. Izhikevich, Int. J. Bifurcat. Chaos 10, 1171 (2000).
[37] H. C. Tuckwell, J. Theor. Biol. 127, 427 (1987).
[38] R. F. Fox and Y. Lu, Phys. Rev. E 49, 3421 (1994).
15
[39] P. Hanggi and H. Thomas, Phys. Rep. 88, 207 (1982).
[40] G. Schmid and P. Hanggi, Math. Biosci. 207, 335 (2007).
[41] P. E. Kloeden, E. Platen,
and H. Schurz, Numerical solution of SDE
through computer experiments (Springer, Berlin, 1994).
[42] G. E. P. Box and M. E. Muller, Ann. Math. Stat. 29, 610 (1958).
[43] H. S. Seung, J. Comput. Neurosci. 9, 171 (2000).
[44] P. A. Tass, Europhys. Lett. 59, 199 (2002).
16
|
1109.4657 | 1 | 1109 | 2011-09-21T21:25:02 | The protein dynamical transition is a pseudogap changeover | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.BM"
] | The emergence of biochemical activities in a protein seem to commence with the onset of atomic mean-square displacements along the protein lattice. The ensuing protein dynamical transition has been discussed extensively in the literature, and often with conflicting conclusions. Here we clarify the phenomenon by establishing a deep connection between the dynamical transition and the pseudogap state where high-temperature superconductivity comes to its end.
For this we first show how to endow proteins with an order parameter akin the quasiparticle wave function in superconductors. We then present universality arguments to claim that the protein dynamical transition takes place in tandem with a pseudogap transmutation. We confirm that available experimental data fully supports our proposal. | physics.bio-ph | physics | The protein dynamical transition is a pseudogap changeover
Andrei Krokhotin1 and Antti J. Niemi2, 1
1Department of Physics and Astronomy, Uppsala University, P.O. Box 803, S-75108, Uppsala, Sweden
2 Laboratoire de Mathematiques et Physique Theorique CNRS UMR 6083,
F´ed´eration Denis Poisson, Universit´e de Tours, Parc de Grandmont, F37200, Tours, France
The emergence of biochemical activities in a protein seem to commence with the onset of atomic
mean-square displacements along the protein lattice. The ensuing protein dynamical transition has
been discussed extensively in the literature, and often with conflicting conclusions. Here we clarify
the phenomenon by establishing a deep connection between the dynamical transition and the pseu-
dogap state where high-temperature superconductivity comes to its end. For this we first show how
to endow proteins with an order parameter akin the quasiparticle wave function in superconductors.
We then present universality arguments to claim that the protein dynamical transition takes place
in tandem with a pseudogap transmutation. We confirm that available experimental data fully
supports our proposal.
The protein dynamical transition was first observed
around 30 years ago in myoglobin [1]-[3].
It is now
claimed to be a common property of all hydrated proteins
[4]-[6]. The transition takes place at temperatures that
are somewhere between 180K-240K. Since this coincides
with the temperature range where proteins quite univer-
sally start to display measurable biochemical activities,
it has been proposed that the protein dynamical transi-
tion is concomitant to the onset of life [4]-[6]. But the
detailed nature of the protein dynamical transition, even
its very existence, remains controversial [7], [8]. The phe-
nomenon, if it indeed exists, appears to closely mimic the
properties of the supercooled water that surrounds the
protein [9], [10]: At temperatures below ∼ 150K a flash-
cooled water under ambient pressure is in an amorphous,
glassy state. Between ∼ 150K-240K there is a no man's
land where the time scale for crystallization is too short
to be analyzed with present day techniques. But there is
evidence of an ice-liquid coexistence in this regime [10].
At around 240K water then enters a crystalline state with
a homogeneous ice nucleation that persists until the liq-
uid phase takes over at 273K.
To a large extent protein folding is known to be driven
by a combination of hydrophobic and hydrophilic effects.
Therefore the properties of a hydrated protein should
mirror those of the adjacent water.
Indeed, there is a
wide consensus that the protein dynamical transition is,
likewise a glass transition, driven by two intertwining
processes that are controlled by a combination of the dif-
ferent properties that water has in the thin hydration
shell and in the surrounding bulk. At an atomary level
the two processes are as follows [4]-[6]: At temperatures
below ∼ 180K the covalent protein lattice is in a state
where the individual atoms are predominantly subject to
local thermal fluctuations. These fluctuations can be in-
terpreted as simple harmonic vibrations of atoms around
their lattice equilibrium positions. The Debye-Waller re-
lation connects the atomic mean square displacements
to the experimental B-factors, that grow linearly at low
temperatures
< x2 >T ≈ B
8π2 (cid:39) a · T + b
(1)
When the protein dynamical transition takes place,
atomic displacements begin to correlate and start cov-
ering much larger length scales. Various collective mo-
tions such as fluctuations between different macromolec-
ular sub-states make the scene, take over and rule the in-
ternal dynamics of the protein lattice all the way to phys-
iological temperatures. This becomes reflected in the B-
factors that should now display a more rapid, maybe even
exponential increase as a function of temperature.
The conventional approach to the protein dynamical
transition is very much concentrated on understanding
the fine structure and detailed properties of supercooled
water [4]-[6]. Here we propose a radical departure from
the paradigm way. Our approach is based on the theoret-
ical concept of universality that already has a firm basis
in the description of phase properties of matter in terms
of a relatively small number of relevant and marginal in-
teractions [11]. We employ a combination of universality
arguments with experimental data analysis to conclude
that the protein dynamical transition must be a pseudo-
gap changeover. We argue that the pertinent pseudogap
state is the one that precedes the Θ-point phase transi-
tion where proteins become denatured and depart from
their biologically active collapsed phase. Furthermore,
since our arguments are entirely based on the concept of
universality, the transition can not be specific to proteins
and water only, it is a more general property of polymers
in bad solvents.
A pseudogap state was originally introduced to explain
aspects of high temperature superconductors [12], [13].
Subsequently it has been found experimentally even in
ferromagnetic metals [14]. Theoretically, a pseudogap
has been proposed to be a participant both in color su-
perconductors and in the chiral phase transition of QCD
[15], [16]. At a theoretical level and in its simplest form,
the pseudogap state is described by a single complex or-
1
1
0
2
p
e
S
1
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
5
6
4
.
9
0
1
1
:
v
i
X
r
a
der parameter φ = ρ exp{iθ}. In the case of superconduc-
tivity this could be the order parameter for the Cooper
pairing of electrons. In the symmetric phase where there
is no superconductivity, the free energy has its minimum
when the modulus < φ >=< ρ > vanishes. The symme-
try becomes broken when there is condensation of φ and
the free energy is minimized by a non-vanishing value of
the modulus < ρ >. In the case of a superconductor this
takes place when electrons combine into Cooper pairs,
the hallmark of a superconducting phase. The pseudo-
gap state is a refinement of this phase structure into the
case where < φ > vanishes even though < ρ > retains its
non-vanishing symmetry breaking value [12]. This occurs
in the presence of a strong phase decoherence,
< exp{iθ} >= 0
(2)
Consequently we can have < φ >= 0 even though the
nonvanishing gap < ρ > persists. The pseudogap state
is like a symmetric phase precursor state in the broken
symmetry phase, a prelude to the transition that takes
place when the gap < ρ > eventually vanishes.
In the case of proteins, more generally of polymers, the
conventional order parameter is the compactness index ν
that describes how the radius of gyration Rg scales in the
limit where the number N of residues becomes very large
[17],
Rg =
(ri − rj)2 ≈ R0N ν
(3)
(cid:115) 1
(cid:88)
2N 2
i,j
Here ri (i = 1, 2, ..., N ) are the locations of the central
Cα carbons and R0 is a protein specific but N indepen-
dent pre-factor. Canonically, the compactness index can
assume only three different values. The mean-field the-
ory value ν = 1/3 corresponds to the biologically active
collapsed phase where proteins are in a space filling con-
formation. This commonly occurs at low temperatures
or in a bad solvent environment. When ν = 1/2 we have
a fully flexible chain, this corresponds to the Θ-point.
Finally, ν = 3/5 is the mean-field Flory value for the
universality class of self-avoiding random walk that de-
scribes a polymer either at a very high temperature or
in a very good solvent environment. This conventional
approach to protein phase structure does not have the
flexibility to describe a pseudogap state, so we now pro-
pose a refinement. We start by introducing the local
backbone bond and torsion angles. We utilize the Cα
coordinates to define the backbone unit tangent (t) and
binormal (b) vectors
ti =
ri+1 − ri
ri+1 − ri & bi =
ti−1 × ti
ti−1 × ti
(4)
Then
κi = arccos(ti+1 · ti) & τi = arccos(bi+1 · bi)
2
are the discrete bond and torsion angles, respectively.
Note that these definitions involve only the positions of
the Cα carbons. Consequently each quantity is inherently
geometrical and we can introduce the similarly geomet-
rical order parameter
ψi = κi · exp{iτi}
(6)
In fact, this is exactly the Hasimoto identification of the
wave function in nonlinear Schodinger equation [18], in
its discrete version. As a complex variable the order pa-
rameter (6) also has the requisite structure to describe a
pseudogap. We now proceed to argue that in a collapsed
protein the pseudogap state also occurs, and that it is a
precursor state to the Θ-point phase transition.
In [19] it has been shown that most proteins in Pro-
tein Data Bank, over 90% of them, can be described with
experimental B-factor accuracy in terms of soliton solu-
tions to the following variant of the discrete nonlinear
Schrodinger equation [20]
E = − N−1(cid:88)
(cid:26)
N(cid:88)
2 κi+1κi +
i + q · (κ2
2κ2
i=1
i=1
+
dτ
2
i − bτ κ2
i τ 2
κ2
i τi − aτ τi +
cτ
2
τ 2
i
i − m2)2
(cid:27)
(7)
The first sum together with the three first terms in the
second sum comprise exactly the energy of the standard
discrete nonlinear Schrodinger equation (DNLS) when
expressed in terms of the Hasimoto variable (6). The
fourth (bτ ) and fifth (aτ ) terms are the two conserved
quantities that precede the energy in the DNLS hierar-
chy, the momentum and the helicity [21]. The last (cτ )
term is the Proca mass term, we include it only for com-
pleteness.
The energy function (7) does not purport to explain
the details of the atomary level mechanisms that give
rise to protein folding. Rather, it allows us to exam-
ine the properties of a folded protein backbone in terms
of universal physical arguments, much like an effective
Landau-Ginzburg model describes superconductivity. In
fact, we can further develop the analogy between (6) and
Cooper pairs, by noting that (7) has the functional form
of the discretized Landau-Ginzburg free energy in the su-
percurrent variables [22]: In the continuum limit the first
two terms combine into the derivative of κ(x) that plays
the role of Cooper pair density. The third term is the
standard symmetry breaking potential and the fourth
term has its origin in spontaneous symmetry breaking
that leads to the notorious Meissner effect [22]. In addi-
tion we have included in (7) exactly all those terms that
are consistent with general principles of universality and
gauge invariance.
(5)
If we use the τi equation of motion to eliminate the
torsion angles, the potential for the bond angles becomes
(cid:18)
(cid:19)
Epot =
ad + bc
d(c + dκ2)
+ 2
1 − qm2 − b2
4d
κ2 + qκ4
(8)
The first term relates to the potential energy in a
Calogero-Moser system, notorious for its role in describ-
ing fractional statistics. Here it is slightly deformed by
the Proca mass. The second and third terms have the
conventional form of a symmetry breaking double-well
potential. Depending on the parameter values we may
be either in the symmetric κi = 0 phase or in the bro-
ken symmetry phase where κi acquires a non-vanishing
ground state value.
In the second case the local en-
ergy minimum states correspond to regular protein sec-
ondary structures such as α-helix or β-strand and pro-
tein loops are like domain wall configurations that inter-
polate between the different local minima [23]. Explicit
computations confirm [24] that the symmetry breaking
ν ≈ 1/3 low temperature phase where long-range corre-
lations rule, becomes converted into a symmetry restored
ν ≈ 1/2 phase at higher temperatures.
In particular,
the symmetry restoration is a phase transition that takes
place when temperature reaches the Θ-point value that
separates the low temperature collapsed phase from the
higher temperature ideal chain phase [24].
Unlike in the case of conventional metallic supercon-
ductors where the supercurrent has stiffness that leaves
no room for a pseudogap, the torsion angles τi in (7) are
only subject to ultra-local couplings and as such they are
highly flexible. Thus it is conceivable that as temperature
starts increasing in the low temperature phase, there will
also be an increase in phase decoherence that is measured
by < exp{iτ} >. Depending on the dynamical details,
the energy function (7) may then describe a pseudogap
state at temperatures below the Θ-point value. But as it
stands, the present energy function can not fully model
the pseudogap transition in proteins. This is because the
model does not include the side-chains with their numer-
ous rotamers χi. The rotamers give us plenty of opportu-
nities to introduce many additional pseudogap detecting
order parameters < exp{iχi} > in proteins:
At low temperatures protein is in a crystalline state
and the rotamers assume very definite values that are cat-
alogued in various libraries. As temperature raises fluc-
tuations appear, but these are damped by various steric
constraints that strongly limit the ability of individual
atoms to move.
In proteins the steric constraints are
more lenient to the outlying side-chain rotamers than to
the backbone torsion angles: With fever covalent neigh-
bors, there is more room to fluctuate. Consequently any
protein pseudogap transition should become primarily
visible in an increased phase de-coherence in the outlying
side-chain rotamers. But even though steric constraints
are more lenient to side-chain rotamers, the identification
of a protein pseudogap is much more delicate than the
identification of a pseudogap state in materials such as
3
high-temperature superconductors where no steric con-
straints exist. Despite being numerous, the protein side-
chain rotamers can never assume unhindered, mutually
fully uncorrelated values. Consequently protein rotamer
angles can never display a full, unrestrained phase de-
coherence the way how it is laid out by a high temper-
ature superconductor. Rather we expect there to be a
pseudogap changeover that makes its presence known in
a marked, rapid relative increase in the side-chain ro-
tamer fluctuations as the temperature climbs up towards
the Θ-point value.
Presently, there are no direct experimental measure-
ments of temperature dependence in the rotamer pseu-
dogap order parameters < exp{iχi} >. We propose that
such experiments could be performed. At the moment
our ability to investigate the inception of a pseudogap is
limited to the analysis of increased temperature depen-
dence in the experimental side-chain B-factors that indi-
rectly measure the < exp{iχi} > fluctuations. We have
screened all data available in Protein Data Bank, but mu-
tually compatible good quality PDB data on B-factors is
sparse. Structures have been determined without stan-
dardized crystalline environments, using different refine-
ment procedures, and with apparently different practices
for determining the B-factors. This makes it very hard
for us to compare data that is collected in different ex-
periments and at different temperatures. Consequently
our ability to reliably investigate the relation between a
pseudogap changeover and the protein dynamical tran-
sition is limited to the re-analysis of data that has been
collected in only two experiments, with crambin [25] and
ribonuclease A [26]. They both measure B-factors at dif-
ferent temperatures, consistently with the same protein
crystals. We have re-analyzed the data and find that both
experiments fully support our proposal that the protein
dynamical transition is a pseudogap changeover.
In Figure 1 we display the B-factors of these two exper-
iments as a function of temperature, separately for the
backbone Cα carbons and for the side-chain atoms. From
the side-chains we exclude the Cβ atoms, they are slaved
to the sp3 hybridized backbone Cα atoms both by cova-
lent bonds and by existing refinement procedures. In the
case of crambin we observe a clearly visible qualitative
change and a definite enhancement in the temperature
dependence of the side-chain B-factors in comparison to
the backbone Cα atoms. There is also a transition from a
linear to a non-linear behavior. The change commences
with the protein dynamical transition regime. For ri-
bonuclease A the results are similar, but the effect is less
profound. We remark that in both cases the B-factors of
the Cα atoms truly deviate only slightly (if at all in the
case of ribonuclease A) from linearity in T . This is con-
sistent with [27], where no sign of a protein dynamical
transition was observed at the level of the backbone Cα
carbons.
In summary, the protein dynamical transition and the
A.J.N. thanks H. Frauenfelder for communications,
and for providing a copy of [8].
4
[1] F. Parak, E.N. Frolov, R.L. Mossbauer and V.I. Goldan-
skii, J. Mol. Biol. 145 825 (1981)
[2] R.H. Austin, K.W. Beeson, L. Eisenstein, H. Frauen-
felder and I.C. Gunsalus, Biochemistry 14 5355 (1975).
[3] W. Doster, S. Cusack and W. Petry, Nature 337 754
(1989).
[4] D. Ringe and G.A. Petsko, Biophys. Chem. 105 667
(2003)
[5] H. Frauenfelder et.al. PNAS 106 5129 (2009)
[6] W. Doster, Biochim. Biophys. Acta 1804 3 (2010)
[7] Y. He, P.I. Ku, J.R. Knab, J.Y. Chen and A.G. Markelz,
Phys. Rev. Lett. 101 178103 (2008)
[8] R.D. Young, H.Frauenfelder and P.W. Fenimore, preprint
LA-UR 11-00161
[9] A. Mishima and H.E. Stanley, Nature 396 329 (1998)
[10] E.B. Moore and V. Molineroa, J. Chem. Phys. 132
244504 (2010)
[11] J. Zinn-Justin, Quantum field theory and critical phe-
nomena (Clarendon Press, Oxford, 2002)
[12] V.J. Emery and S.A. Kivelson, Nature 374, 434 (1995).
[13] J. Corson, R. Mallozzi, J. Orenstein, J.N. Eckstein and
I. Bozovic, Nature 398, 221 (1999)
[14] N. Mannella et.al. Nature 438, 474 (2005)
[15] H. Kleinert and E. Babaev, Phys. Lett. B438 311 (1998)
[16] K. Zarembo, JETP Letters 75 59 (2002)
[17] P.G. De Gennes, Scaling Concepts in Polymer Physics
(Cornell University Press, Ithaca, 1979)
[18] H. Hasimoto, J. Fluid Mech. 51 477 (1972)
[19] A. Krokhotin, A.J. Niemi and X. Peng,
e-print
arXiv:1109.3903v1 [physics.bio-ph]
[20] N. Molkenthin, S. Hu and A.J. Niemi, Phys. Rev. Lett.
106 078102 (2011)
[21] L.D. Faddeev and L.A. Takhtajan, Hamiltonian methods
in the theory of solitons (Springer Verlag, Berlin, 1987)
[22] P.G. De Gennes, Superconductivity of Metals and Alloys
(Westfield Press, New York 1995)
[23] M. Chernodub, S. Hu and A.J. Niemi, Phys. Rev. E82
011916 (2010)
[24] M. Chernodub, M. Lundgren and A.J. Niemi, Phys. Rev.
E83 011126 (2011)
[25] M.M. Teeter, A. Yamano, B. Stec and U. Mohanty, Re-
lation to protein function PNAS 98 11242 (2001)
[26] R.F. Tilton, J.C. Dewan and G.A. Petsko, Biochemistry
31 2469 (1992)
[27] A. Krokhotin, A.J. Niemi and X. Peng, (to appear)
FIG. 1: a) The B-factors of crambin from [25] as a function of
temperature, separately for backbone Cα (red) and side-chain
atoms (black) but excluding the Cβ. b) Same as a) but for
ribonuclease A, with data from [26]. In both cases the side-
chains display a clear transition and nonlinear T dependence,
while for the Cα even the linear fit (1) is feasible.
pseudogap state in high temperature superconductors are
two a priori totally different physical phenomena. How-
ever, here we have proposed that they are intimately re-
lated by asserting that the protein dynamical transition
is a pseudogap changeover. Our argumentation is based
both on theoretical observations that stem from the con-
cept of universality, and analysis of all presently available
experimental data. We propose that experiments could
be performed to directly detect the pseudogap transition,
in particular since the possibility that a pseudogap con-
verts a static protein crystal into a biologically active
nano-engine could have really far reaching consequences
in the search for a physical origin of life.
Temperature (K)100150200250300)2ÅTemperature factor (0246810aTemperature (K)100150200250300)2ÅTemperature factor (0510152025b |
1804.00979 | 2 | 1804 | 2019-03-26T05:28:03 | Non-minimum phase viscoelastic properties of soft biological tissues | [
"physics.bio-ph"
] | Understanding the visocoelastic properties of soft biological tissues is important for progress in the field of human healthcare. This study analyzes the viscoelastic properties of soft biological tissues using a fractional dynamics model. We conducted a dynamic viscoelastic test on several porcine samples, namely liver, breast, and skeletal muscle tissues, using a plate--plate rheometer. We found that some soft biological tissues have non-minimum phase properties; that is, the relationship between compliance and phase delay is not uniquely related to the non-integer derivative order in the fractional dynamics model. The experimental results show that the actual phase delay is larger than that estimated from compliance. We propose a fractional dynamics model with the fractional Hilbert transform to represent these non-minimum phase properties. The model and experimental results were highly correlated in terms of compliance and phase diagrams and complex mechanical impedance. We also show that the amount of additional phase delay, defined as the increase in actual phase delay compared to that estimated from compliance, differs with tissue type. | physics.bio-ph | physics |
Non-minimum phase viscoelastic properties of soft biological tissues
Yo Kobayashi∗
Graduate School of Engineering Science,
Osaka University, Osaka, Japan
and JST-PRESTO
Faculty of Science and Engineering / Future Robotics Organization, Waseda University, Tokyo, Japan
Naomi Okamura, Mariko Tsukune, and Masakatsu G. Fujie
Graduate School of Engineering Science, Osaka University, Osaka, Japan
(Dated: March 27, 2019)
Masao Tanaka
Understanding the visocoelastic properties of soft biological tissues is important for progress in
the field of human healthcare. This study analyzes the viscoelastic properties of soft biological
tissues using a fractional dynamics model. We conducted a dynamic viscoelastic test on several
porcine samples, namely liver, breast, and skeletal muscle tissues, using a plate -- plate rheometer. We
found that some soft biological tissues have non-minimum phase properties; that is, the relationship
between compliance and phase delay is not uniquely related to the non-integer derivative order in
the fractional dynamics model. The experimental results show that the actual phase delay is larger
than that estimated from compliance. We propose a fractional dynamics model with the fractional
Hilbert transform to represent these non-minimum phase properties. The model and experimental
results were highly correlated in terms of compliance and phase diagrams and complex mechanical
impedance. We also show that the amount of additional phase delay, defined as the increase in
actual phase delay compared to that estimated from compliance, differs with tissue type.
I.
INTRODUCTION
B. Related research
A. Background
Understanding the physical phenomena in the hu-
man body is important in bioscience and bioengineering.
Knowledge of the mechanical properties of human tissues
will lead to progress in healthcare. In particular, under-
standing the viscoelastic properties of biological tissues is
key because they can reveal tissue function. These prop-
erties are also important for medical treatment because
they are closely related to tissue type and disease.
Nevertheless, methods for analyzing the viscoelastic
properties of soft biological tissues are not well estab-
lished. The properties of soft biological tissues are differ-
ent from those of synthetic materials and thus cannot be
directly modeled in the same manner [1, 2].
The motivation behind this study is to determine the
viscoelastic properties of soft biological tissues by model-
ing their macroscopic properties. Ideally, a model should
be strongly correlated with the experimental data and
have a small number of parameters. A small number of
model parameters is important for determining the vis-
coelasticity of soft biological tissues, the identification of
tissue function, and the robust discrimination of tissue
type based on viscoelasticity.
∗ [email protected]
Many studies have reported that soft biological tis-
sues have viscoelastic properties [1 -- 3]. An ordinary dif-
ferential equation, such as that in the Voigt, Maxwell,
or Kelvin model, is generally used to model viscoelas-
tic properties[1 -- 4]. Models with a small-order ordinary
differential equation do not well fit experimental data
for biological tissues. A large-order ordinary differential
equation such as that in the generalized Maxwell model
can be used to increase model accuracy at the cost of a
large number of model parameters. For example, studies
have modeled the nonlinear viscoelasticity of the brain
[5, 6], kidney [7], breast [8], liver [7, 9 -- 13], skeletal mus-
cle [14], and subcutaneous tissue [15].
Fractional differential equations have recently been
shown to be efficient in modeling the viscoelastic proper-
ties of biological tissues. The fractional dynamics model
represents a power law response, which is obtained from
experimental data of soft biological tissues, with a rela-
tively small number of parameters [16, 17]. For exam-
ple, the fractional dynamics model was used to model
the viscoelastic properties of the lung [18 -- 20], the brain
[21, 22], skeletal muscle [22 -- 24], tendons [25], cultured
cartilage tissues [26], and cells [26 -- 28]. Fractional mod-
els such as the springpot model have been used to analyze
the response in research on magnetic resonance elastog-
raphy [22, 23]. Fractional dynamics has become popular
for modeling viscoelasticity, with experimental data and
models reported for vessels [16], the lung [18 -- 20], skeletal
muscle [22 -- 24, 29, 30], the brain [21, 22], tendons [25],
the liver [22, 31 -- 35], breast tissues [36, 37], muscle cells
[26], blood cells [27], and living cells [28].
The fractional dynamics model has been applied to a
wide variety of materials, including biological materials.
We previously developed a viscoelastic model based on
the fractional dynamics model [29 -- 37]. The model was
derived using experimental data obtained from in vitro
measurements of a porcine liver [31 -- 34]. We also vali-
dated the model with data obtained for in vitro breast
tissue (mammary gland, fat, and muscle) [36, 37]. The
model was partially evaluated using data for in vitro and
in vivo skeletal muscle tissue [29, 30].
Our previous study [35] also investigated the dynamic
viscoelastic properties of liver tissue and evaluated the
pairing of compliance J(ω) and phase delay φ(ω). The
study showed that liver tissue has a power-law decrease
in compliance J and a constant phase delay φ in the fre-
quency domain. These characteristics can be accurately
represented using a fractional dynamics model.
In the
experiment and model, the compliance J and phase de-
lay φ were found to be causally related via a non-integer
derivative order α, specifically J ∝ ω−α, φ = − π
2 α. In
this case, the dynamic viscoelastic properties of liver tis-
sue are represented by minimum phase properties [38]. In
previous studies, we also conducted a dynamic viscoelas-
tic experiment on breast [36, 37] and muscle [29] tissues.
The results showed that the experimental data of breast
and muscle tissues are not as highly correlated with the
fractional model as the data for liver tissue are.
C. Objectives
The objective of this study is to develop a fractional
dynamics model that represents the viscoelastic proper-
ties of soft biological tissues. Specifically, a dynamic vis-
coelasticity test, which gives the frequency response, was
conducted. We found that skeletal muscle and breast
tissues have non-minimum phase properties; that is, the
relationship between compliance and phase delay is not
uniquely related to a non-integer derivative order α. The
experimental results show that the actual phase delay is
larger than that estimated from compliance.
This paper proposes a model for representing the non-
minimum phase properties obtained from a dynamic vis-
coelasticity test. We also show the amount of additional
phase delay, defined as the increase in actual phase delay
compared to that estimated from compliance, for several
tissue types. Figure 1 shows an overview of this article.
II. MATERIALS AND METHODS
A. Materials
We investigated the viscoelasticity of several types of
porcine tissue, namely the liver, mammary gland, breast
muscle, breast fat, and psoas major muscle, longissimus
2
thoracis muscle, and muscle fat. For the liver [35] and
breast [36] tissues, we used the experimental data from
a dynamic viscoelastic test reported in a previous study.
We conducted an experiment on skeletal muscle tissues.
Figure 1 (a-2) shows the details of the measurement
setup.
B. Experimental setup and procedure
The experimental setup and procedure are almost the
same as those described in a previous article [35]. A
description is given in this section to enhance the read-
ability of this article.
We used a plate -- plate rheometer (AR-G2 or DHR2;
TA Instruments, New Castle, DE) to measure the stress
and strain of the sample. A shear stress rheometer was
selected because the shear test must be independent of
any change in the cross-sectional area in the stress calcu-
lation. In addition, with this device, the effect of gravity
can be disregarded. From these measurements, the con-
ventional shear strain x and conventional shear stress f
were calculated. The measurements of strain x and stress
f are valid only when there is no slip between the sample
and the plates. Thus, sandpaper was attached to the top
plate and the measurement table to prevent sliding. The
samples were cut into slices (diameter: 20 mm; thickness:
about 5 mm), which were placed on a measurement ta-
ble. The samples were soaked in a saline solution at 35◦C
during testing.
After the saline solution had reached the target tem-
perature, the gap between the table and the top plate
was zeroed to the surface of the saucer. The saline solu-
tion was stable, and there was no reflux flow. Each tis-
sue sample was placed on a measurement table, and the
sample thickness (i.e., gap) was determined. The sample
thickness was defined as the distance between the sur-
face of the saucer and the surface of the parallel plate
(part of the measurement device) at the time that the
normal stress resulting from the contact between the par-
allel plate and the sample reached 0.1 N. To engage the
sample and parallel plate, preloading for over 100 sec-
onds and unloading for over 100 seconds were performed
three times under a constant shear stress of 375 Pa. The
following series of experiments were conducted for each
sample after the above initialization procedures.
A sine-wave stress of 0.1 to 10 rad/s, providing a 1.5%
strain amplitude, was applied to the sample. The strain
amplitude of 1.5% (= 0.015) is within the range in which
all tissues exhibited linear responses. The compliance J,
phase delay φ, storage elastic modulus G', and loss elastic
modulus G" at various angular frequencies ω were mea-
sured. Details of the process used to obtain the experi-
mental results from the dynamic viscoelastic test are de-
scribed in [35]. The effects of the mass (inertia) and shear
viscosity of the external normal saline solution could be
disregarded at frequencies of lower than 10 rad/s. Data
were collected for each tissue type. The number of sam-
3
FIG. 1. Visual overview of this article. The viscoelastic properties were investigated from material measurements of several
biological tissues. We used porcine tissues (liver, mammary gland, breast muscle, breast fat, psoas major muscle, longissimus
thoracis muscle, and muscle fat) as samples (a-1). We used a plate -- plate rheometer, which can dynamically control and measure
the stress and strain applied to the sample, to measure the samples (a-2). We conducted a dynamic viscoelastic test (b) to
measure the viscoelastic properties and derive the viscoelastic model (c). We found that skeletal muscle and breast tissues
have non-minimum phase properties; that is, the relationship between compliance and phase delay is not uniquely related to
the fractional derivative order. The experimental results show that the actual phase delay is larger than that estimated from
compliance. This paper also proposes a model for representing the non-minimum phase properties.
ples for each tissue type is shown in Table I. We obtained
pairs of results, (compliance J, phase φ) or (storage elas-
tic modulus G', loss elastic modulus G", from the dy-
namic viscoelastic test.
III. RESULTS AND MODELING
A. Compliance and phase delay
Typical experimental results of the compliance and
phase of a sample for each tissue type are shown in Fig.
2, where compliance J is the multiplicative inverse of G∗.
The experimental data for all samples of a given tissue
exhibited the same trend as that of the typical sample.
The power-law compliance J decreases as the angular fre-
quency ω increases for over two decades. The phase delay
φ remains constant as the angular frequency ω changes
for over two decades.
The liver tissue response in the log-log diagram shown
in Fig. 2 (a) has almost the same slope as those for the
mammary gland, breast muscle, breast fat, and psoas
major muscle, shown in Fig. 2 (b)-(e), respectively. This
means that the power law index α from the compliance
data is almost the same among these tissues. The phase
delay in breast muscle, breast fat, and psoas major mus-
cle is larger than that in liver tissue. The slope for longis-
simus thoracis muscle and muscle fat response, shown in
Fig. 2 (f)-(g) differs from the other tissues. A model used
in previous research on liver tissue [35] showed that the
relationship between compliance (J ∝ ω−α) and phase
delay (φ == − π
2 α) is uniquely related to the derivative
order α. The response of the liver tissue almost satisfies
this relationship, but those of the other tissues do not.
Thus, we found that some soft biological tissues have an
additional phase delay, namely, the difference between
the experimentally measured phase delay and the phase
delay (φ = − π
2 α) estimated from the compliance data
(J ∝ ω−α) .
Here, we introduce a model that represents the charac-
teristics of the experimental results, including the power-
law form of compliance, constant phase delay, and addi-
tional phase delay. Our model is given in equation (1).
Equation (2) is a model introduced in a previous article
[35]; it is used as a reference.
H∆αtα
r
dα
dtα (Gx) = f
tα
r
dα
dtα (Gx) = f
(1)
(2)
where x is the strain (torsional strain), f is the stress
(torsional stress), t is time, α is a non-integer derivative
order representing the viscoelasticity ratio, tr is the refer-
ence time scale, G is the linear viscoelastic stiffness at an
arbitrarily chosen point in time tr, H∆α is the fractional
Hilbert transform operator of the order ∆α [39], and ∆α
is an additional phase delay ratio used to represent the
non-minimum amount of the system (i.e., an index of the
additional phase delay). The term H∆α is the fractional
Hilbert transform operator, which is used to represent an
additional phase delay. Equation (1) is equal to equation
(2), presented in our previous study, when ∆α = 0 [39].
The equation is expanded below to explain the above
characteristics. The frequency transfer function is:
J(jω) =
X(jω)
F (jω)
=
1
j∆α tα
r G(jω)α =
(cid:16)
1
j∆αG
j ω
ωr
(cid:17)α (3)
4
FIG. 2. Compliance and phase delay diagrams. Typical experimental results for a sample of each tissue type are shown.
Results for (a) liver, (b) mammary gland, (c) breast muscle, (d) breast fat, (e) psoas major muscle (fillet), (f) longissimus
thoracis muscle (loin), and (g) muscle fat. The plots show experimental data. All samples for a given tissue type exhibit the
same trend as that of the typical sample. The power-law compliance J decreases as the angular frequency ω increases for
over two decades. The phase delay φ remains constant as the angular frequency ω changes for over two decades. The lines
show the compliance J and phase φ of our model in equations (5) and (6). The solid line shows the compliance J. The dotted
line shows the phase delay π
2 (α + ∆α). The
difference between the dashed line and dotted line in the phase model is π
2 ∆α. For the liver tissue, there are a few differences
between the dashed and dotted lines. Thus, the liver tissue has minimum phase viscoelastic properties. The other tissues have
non-minimum phase viscoelastic properties. The model, which was fit to the typical experimental data through parameter
identification, shows that the data of our model and the experimental data are highly correlated.
2 α estimated from the compliance data. The dashed line shows the phase delay π
Here, ω is the angular frequency, j is the imaginary
unit, and ωr is the reference scale, which is defined as
ωr = 1/tr. We use the following relationship: H∆α =
exp(j π
2 ∆α) = j∆α [39].
The compliance J is defined from equation (3) as fol-
lows:
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
J(ω) =
1
j∆αG(j ω
ωr
)α
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) =
(cid:16) ω
1
(cid:17)α =
(cid:16) ω
(cid:17)α
J(ωr)
ωr
ωr
G
where J(ωr) is a coefficient representing compliance,
which is defined as J(ωr) = 1/G.
Equation (5) is derived from the log-log transforma-
tion of (4) through a transformation into dimensionless
follows:
(4)
quantities.
(cid:18) J(ω)
(cid:19)
J(ωr)
log
(cid:18) ω
(cid:19)
ωr
= − α log
(5)
The model equation of the phase delay φ is derived as
(cid:18)
1
φ (ω) = arg
= − arg(cid:0)j∆α(cid:1) − arg
j∆αG(j ω
(cid:16)
2 α = − π
2 ∆α − π
= − π
= φo
(cid:19)
(cid:16)
ωr )α
G
j ω
ωr
2 (α + ∆α)
(cid:17)α(cid:17)
(6)
where φo (= − π
resents the phase delay.
2 (α + ∆α)) is the coefficient that rep-
Thus, our model represents the trends in the experi-
mental results, namely the decrease in power-law compli-
ance, as (5), constant phase delay, and additional phase
delay, as (6).
We fitted the compliance J(ω) and phase φ(ω) of our
model to the experimental results through parameter
identification. Specifically, the parameters G, α, and ∆α
were identified for each sample. Details of the method,
process, and equation used in the parameter identifica-
tion are provided in Appendix.
The data of compliance J(ω) and phase φ(ω) from our
model, in which the parameters were fitted to the ex-
perimental data, are shown in Fig. 2. The figure shows
that the data of our model and the experimental data are
strongly correlated. Table I lists the fundamental statis-
tics about the model parameters for each tissue type. In
Table I, each dataset for a single experiment was fitted
to identify the set (G, α, ∆α) of model parameters. The
results of these parameters were then averaged.
B. Mechanical impedance
In this section, we present the results of mechanical
impedance. The mechanical complex impedance G∗ is
defined as follows:
G∗(ω) = G(cid:48)(ω) + jG(cid:48)(cid:48)(ω)
(7)
Here, ω is the angular frequency, G∗ is the complex
mechanical impedance, G' is the storage elastic modulus,
and G" is the loss elastic modulus.
Typical experimental data of the mechanical complex
impedance G∗ for a sample of each tissue type are shown
in Fig. 3. This figure was made using the data in Fig.
2. All samples for each tissue type exhibited the same
trend as that of the typical sample. The storage elas-
tic modulus G(cid:48) and the loss elastic modulus G(cid:48)(cid:48) increase
with increasing angular frequency ω. The data of G(cid:48) and
G(cid:48)(cid:48) exhibit a power-law form for over two decades. The
slopes of G(cid:48) and G(cid:48)(cid:48) in the log-log diagram are almost the
same.
Our model shows the same characteristics as those of
the experimental data, such as the power-law forms of G(cid:48)
and G(cid:48)(cid:48) with the same slopes. The equation is expanded
below to explain the above results. Because equation
(1) takes the form of a frequency transfer function, the
complex shear modulus G∗ can be expressed as follows:
G∗(jω) = F (jω)
= j∆αG
X(jω)
j ω
ωr
(cid:17)α
(cid:16) ω
ωr
= G
(cid:17)α
j(α+∆α)
(8)
(cid:16)
separation of the real and imaginary parts of (8).
G(cid:48)(ω) = G(cid:48)(ωr)
G(cid:48)(cid:48)(ω) = G(cid:48)(cid:48)(ωr)
(cid:19)α
(cid:19)α
(cid:18) ω
(cid:18) ω
ωr
ωr
Here, G(cid:48)(ωr) and G(cid:48)(cid:48)(ωr) are constant parameters that
represent the storage elastic modulus and the loss elastic
modulus, respectively. The parameters have the follow-
ing relationship (10):
(cid:113)
G =
G(cid:48)(ωr)2 + G(cid:48)(cid:48)(ωr)2
G(cid:48)(ωr) = G cos(
G(cid:48)(cid:48)(ωr) = G sin(
π
2
π
2
(α + ∆α))
(α + ∆α))
(cid:18) G(cid:48)(ω)
(cid:18) G(cid:48)(cid:48)(ω)
G(cid:48)(ωr)
(cid:19)
(cid:19)
G(cid:48)(cid:48)(ωr)
log
log
(cid:18) ω
(cid:18) ω
ωr
(cid:19)
(cid:19)
ωr
= α log
= α log
5
(9a)
(9b)
(10a)
(10b)
(10c)
(11a)
(11b)
Equations (11a) and (11b) were derived from (9a) and
(9b) using a log-log transformation through a transfor-
mation into dimensionless quantities.
Thus, our model equation exhibits the same trend as
that of the experimental data, i.e., the power-law depen-
dence of the storage elastic modulus G(cid:48) and the loss elas-
tic modulus G(cid:48)(cid:48). The additional phase parameter ∆α
affects the ratio of storage elastic modulus G(cid:48)(wr) to
loss elastic modulus G(cid:48)(cid:48)(wr). This ratio for the model
without the additional phase term (2) is related to the
power law index α as G(cid:48)(wr)/G(cid:48)(cid:48)(wr) = tan( π
2 α). This
ratio for the model with the additional phase term (1) is
G(cid:48)(wr)/G(cid:48)(cid:48)(wr) = tan( π
2 (α + ∆α)).
The parameters G, α, and ∆α were identified by fit-
ting the experimental data for all samples of each tissue
type. The G' and G" in our model, which fit the typical
experimental data, are presented in Fig. 3. This figure
shows that the data of our model and the experimental
data are strongly correlated. The coefficient of determi-
nation R2 between our model and the experimental data
for the series of G' and G" for all samples of each tissue
type is approximately 90%. Table I lists the fundamental
statistics of the model parameters for each tissue type. In
the table, each dataset for a single experiment was fitted
to identify the set (G, α, and ∆α) of model parameters.
The results of the model parameters were then averaged.
IV. DISCUSSION
Here, we use the following relationship: H∆α = j∆α [39].
Equation (8) expands to (9a) and (9b) from (7) with a
The main contribution of this article is the identifica-
tion of the non-minimum phase viscoelastic properties of
6
FIG. 3. Mechanical complex impedance. Typical experimental data for the samples of each tissue type are shown. The
plus and cross symbol plots respectively show the experimental data for the storage elastic modulus G(cid:48) and the loss elastic
modulus G(cid:48)(cid:48). Results for (a) liver, (b) mammary gland, (c) breast muscle, (d) breast fat, (e) psoas major muscle (fillet), (f)
longissimus thoracis muscle (loin), and (g) muscle fat. All samples for a given tissue type exhibit the same trend as that of
the typical sample. G(cid:48) and G(cid:48)(cid:48) increase as the angular frequency ω increases. Both G(cid:48) and G(cid:48)(cid:48) exhibit a power-law form for
over two decades. The slopes of G(cid:48) and G(cid:48)(cid:48) in the log-log diagram are almost the same. The data of G(cid:48) and G(cid:48)(cid:48) in our model
are indicated by the solid and dashed lines, respectively. These data, which were fit to the typical experimental data through
parameter identification, show that our model and the experimental data are highly correlated.
TABLE I. Fundamental statistics of the model parameters for tr = 1 (ωr = 1)
tissue type
liver
breast gland
breast muscle
breast fat
psoas major muscle
longissimus thoracis muscle
muscle fat
sample
number
G
(Avg.)
6
10
5
12
10
10
10
402
252
753
375
3586
2738
2006
.
G
(S.D.)
132
63
172
149
576
462
471
α
(Avg.)
0.120
0.111
0.116
0.107
0.114
0.097
0.063
α
(S.D.)
0.008
0.010
0.003
0.007
0.011
0.009
0.007
∆α
(Avg.)
0.003
0.042
0.068
0.067
0.092
0.136
0.146
∆α
(S.D.)
0.011
0.008
0.013
0.017
0.014
0.011
0.019
R2
(Avg.)
0.90
0.91
0.91
0.92
0.93
0.94
0.94
soft biological tissues and the development of a model
that represents these properties. Here, minimum phase
systems are defined as systems that have the minimum
phase delay for a given magnitude (compliance in this ar-
ticle) of the response. A minimum phase system has the
smallest possible phase for a give magnitude response.
A system has minimum phase properties when it and
its inverse are causal and stable. A non-minimum phase
system has a phase delay that is larger than that of a
minimum phase system with the equivalent magnitude.
For a fractional-order system with index α, the system
has minimum phase properties when φ = π
2 α, and the
system has non-minimum phase properties when φ > π
2 α
[38]. For a minimum phase system, the relationship be-
tween the magnitude (compliance) and phase delay is
uniquely determined by Bode's theorem, which means
that a phase diagram can be estimated from a magni-
tude diagram, and vice versa. For such a system, the
time response can also be estimated from magnitude and
phase diagrams through the inverse Fourier transform.
The index α in the fractional model for viscoelasticity
is important for characterizing model properties. The
value α can be estimated from several types of experi-
mental data, such as a decrease in power-law compliance
and constant phase delay. The estimation is not limited
to the frequency domain. The time response, such as
the power-law strain increase in the creep test and the
power-law decrease in the stress relaxation test, can also
be used. In this investigation, it was expected that the
same value of α could be obtained in each experiment
under the assumption that the above relationship in a
minimum phase system is satisfied. The results obtained
here show that the index α should be evaluated under
the consideration that soft biological tissues have non-
minimum phase viscoelastic properties. For example, we
found a difference in the estimated index between the de-
crease in power-law compliance (α) and constant phase
delay (α + ∆α).
From a practical point of view, the contribution of this
study is a parameter that is useful for discriminating tis-
sue types. The additional phase delay parameter ∆α dif-
fers with tissue type. In particular, muscle tissues such
as the psoas major muscle, longissimus thoracis muscle,
and muscle fat have very different ∆α values. The ∆α
value may be related to the fat cell content in tissue.
The liver has a simple cell structure and consists mainly
of liver cells. Porcine liver tissue includes only a few fat
cells, whereas breast and skeletal muscle tissues include
many fat cells. In particular, the fat cell content in mus-
cle tissue increases in the order of psoas major muscle,
longissimus thoracis muscle, and muscle fat. The ∆α
value increases in the same order.
The main limitation of this study is that it does not
explain how non-minimum phase properties come about.
The Hilbert transform operator is used in the Benjamin-
Ono equation for internal waves in stratified fluids, where
it is introduced as a theoretical expansion of the phys-
ical model [40, 41]. Further theoretical investigation is
needed regarding the fractional Hilbert transform and
non-minimum phase properties. In addition, the effects
of non-minimum phase properties on the response in the
time domain should be investigated. Finally, the frac-
tional model was partially explained through a fractal
structure in related studies [35, 42]. The actual structure
and how the non-minimum phase viscoelasticity and the
fractional Hilbert transform operator can be related to
the structure are still unknown.
V. CONCLUSION
7
This study proposed a model that represents the
viscoelastic properties of soft biological tissues. We
found that breast and skeletal muscle tissues have non-
minimum phase properties in a dynamic viscoelastic test.
The experimental results show that the actual phase de-
lay is larger than the phase delay π
2 α estimated from
the index α of the power-law compliance. The proposed
model and the experimental results were highly corre-
lated in terms of the compliance and phase diagrams
and the complex mechanical impedance. The additional
phase delay parameter ∆α may be useful for discriminat-
ing tissue types because it differs with tissue type.
ACKNOWLEDGMENTS
This work was supported in part by the Japan Science
and Technology Agency (JST) Precursory Research for
Embryonic Science and Technology (PRESTO) (No. JP-
MJPR14D3), Japan, the Global Centers of Excellence
(GCOE) Program and Grants for Excellent Graduate
Schools, Japan, and a Grant-in-Aid for Scientific Re-
search from the Ministry of Education, Culture, Sports,
Science and Technology (MEXT) (No. 25350577), Japan.
Appendix A: Extended Kalman filter for dynamic
viscoelastic test
The parameter identification method was almost the
same as that described in a previous article [35]. A de-
scription is given in this section to enhance the readabil-
ity of this article. This section shows the methodology
used to identify the parameters described in Sec.
III.
The model for the dynamic viscoelastic test was as fol-
lows, derived from equations (11a) -- (11b):
log
= α log
(cid:18) G(cid:48)(ω)
(cid:18) G(cid:48)(cid:48)(ω)
G(cid:48)(ωr)
(cid:19)
(cid:19)
G(cid:48)(cid:48)(ωr)
log
G(cid:48)(ωr) = G cos(
G(cid:48)(cid:48)(ωr) = G sin(
(cid:18) ω
(cid:18) ω
ωr
(cid:19)
(cid:19)
= α log
ωr
π
2
π
2
(α + ∆α))
(α + ∆α))
(A1a)
(A1b)
(A1c)
(A1d)
where G', G", and ω are variables, and G, α, and ∆α
are parameters.
We obtained the set of G' and G" at each angular
frequency ω value from the experiment. We identified the
parameterfrom these data using the extended Kalman
filter (EKF) (ref.
[43]). System identification using the
EKF can be generally described as follows:
(A2a) and (A2b) are represented as:
θk+1 = Iθk
yk = h(θk) + ζk
8
(A3a)
(A3b)
θk+1 = f (θk, ψk)
yk = g(θk, ζk)
(A2a)
(A2b)
where k = 0, 1, 2,...
represents the discrete itera-
tion index (number of datasets in this case), θ is an n-
dimensional state vector, ψ is an n-dimensional system
noise vector, y is a p-dimensional observation vector, ζ
is a p-dimensional observation noise vector, and f() and
g() are nonlinear vector functions. In state-space theory,
(A2a) and (A2b) are known as the system model (or state
model) and the observation model, respectively.
The parameter vector is regarded as a state vector in
the EKF for system identification. The state vector (pa-
rameter vector) θ is a constant vector and the observation
noise vector ζ is a Gaussian white noise with zero mean.
where I is the identity matrix and h() is a nonlinear vec-
tor function. For system identification for the dynamic
viscoelastic test, the state vector (parameter vector) θ,
observation vector y, and nonlinear vector function h()
are regarded as follows for ωr = 1:
G
(cid:20) log G(cid:48)
(cid:20) α log ω + log(G cos( π
log G(cid:48)(cid:48)
α log ω + log(G sin( π
α
∆α
(cid:21)
θ =
y =
h(θ) =
(A4a)
(cid:21)
(A4b)
(A4c)
2 (α + ∆α)))
2 (α + ∆α)))
The EKF algorithm (ref. [43]) using (A4a) -- (A4c) was
applied to identify the parameter from the dataset.
It
was not necessary to set initial values for each parameter
θ0, meaning that θ0 was a zero vector.
[1] Y. Fung, Biomechanics: mechanical properties of living
tissues, Biomechanics / Y. C. Fung (Springer-Verlag,
1981).
[2] Y.-C. Fung, Biomechanics: mechanical properties of liv-
ing tissues (Springer Science & Business Media, 2013).
[3] W. Maurel, Y. Wu, N. M. Thalmann,
and D. Thal-
mann, Biomechanical models for soft tissue simulation
(Springer, 1998).
[4] A. Wineman, Mathematics and Mechanics of Solids 14,
300 (2009).
and D. Laurendeau, Medical Image Analysis 9, 103
(2005).
[14] B. B. Wheatley, R. B. Pietsch, T. L. H. Donahue, and
L. N. Williams, Computer Methods in Biomechanics and
Biomedical Engineering 19, 1181 (2016).
[15] S. K. Panda and M. L. Buist, Journal of biomechanics
69, 121 (2018).
[16] D. O. Craiem and R. L. Armentano, Engineering in
Medicine and Biology Society, Annual International Con-
ference of the IEEE , 1098 (2006).
[5] K. K. Darvish and J. R. Crandall, Medical engineering
[17] D. Craiem and R. L. Magin, Physical biology 7, 13001
& physics 23, 633 (2001).
(2010).
[6] K. Miller and K. Chinzei, Journal of Biomechanics 35,
483 (2002).
[7] J. Kim and M. a. Srinivasan, Medical image computing
and computer-assisted intervention 8, 599 (2005).
[8] S. Qiu, X. Zhao, J. Chen, J. Zeng, S. Chen, L. Chen,
Y. Meng, B. Liu, H. Shan, M. Gao, and Y. Feng, Journal
of Biomechanics 69, 81 (2018).
[9] S. Marchesseau, T. Heimann, S. Chatelin, R. Willinger,
and H. Delingette, Progress in Biophysics and Molecular
Biology 103, 185 (2010).
[10] B. Ahn and J. Kim, Medical Image Analysis 14, 138
(2010).
[18] B. Suki, a. L. Barab´asi, and K. R. Lutchen, Journal
of applied physiology (Bethesda, Md. : 1985) 76, 2749
(1994).
[19] H. Yuan, E. P. Ingenito, and B. Suki, Journal of applied
physiology (Bethesda, Md. : 1985) 83, 1420 (1997).
[20] H. Yuan, S. Kononov, F. S. Cavalcante, K. R. Lutchen,
E. P. Ingenito, and B. Suki, Journal of applied physiology
(Bethesda, Md. : 1985) 89, 3 (2000).
[21] I. Sack, B. Beierbach, J. Wuerfel, D. Klatt, U. Hamhaber,
S. Papazoglou, P. Martus, and J. Braun, NeuroImage 46,
652 (2009).
[22] I. Sack, K. Johrens, J. Wurfel, and J. Braun, Soft Matter
[11] E. Samur, M. Sedef, C. Basdogan, L. Avtan, and O. Duz-
9, 5672 (2013).
gun, Medical Image Analysis 11, 361 (2007).
[23] D. Klatt, S. Papazoglou, J. Braun, and I. Sack, Physics
[12] P. Asbach, D. Klatt, U. Hamhaber, J. Braun, R. Soma-
sundaram, B. Hamm, and I. Sack, Magnetic Resonance
in Medicine 60, 373 (2008).
[13] J.-M. Schwartz, M. Denninger, D. Rancourt, C. Moisan,
in medicine and biology 55, 6445 (2010).
[24] N. Grahovac and M. igi, Computers and Mathematics
with Applications 59, 1695 (2010), fractional Differenti-
ation and Its Applications.
9
[25] V. D. Djordjevi´c, J. Jari´c, B. Fabry, J. J. Fredberg, and
D. Stamenovi´c, Annals of Biomedical Engineering 31,
692 (2003).
[26] Q. Chen, B. Suki, and K.-N. An, Journal of biomechan-
ical engineering 126, 666 (2004).
[34] Y. Kobayashi, H. Watanabe, T. Hoshi, K. Kawamura,
and M. G. Fujie, in Soft Tissue Biomechanical Modeling
for Computer Assisted Surgery (Springer, 2012) pp. 41 --
67.
[35] Y. Kobayashi, M. Tsukune, T. Miyashita, and M. G.
[27] S. E. Duenwald, R. Vanderby, and R. S. Lakes, Annals
Fujie, Phys. Rev. E 95, 022418 (2017).
of Biomedical Engineering 37, 1131 (2009).
[28] M. Balland, N. Desprat, D. Icard, S. F´er´eol, A. Asnacios,
J. Browaeys, S. H´enon, and F. m. c. Gallet, Phys. Rev.
E 74, 021911 (2006).
[29] Y. Kobayashi, T. Watanabe, M. Seki, T. Ando,
M. G. Fujie, Advanced Robotics 26, 1253 (2012).
and
[30] N. Okamura, M. Tsukune, Y. Kobayashi, and M. G. Fu-
jie, in Engineering in Medicine and Biology Society, An-
nual International Conference of the IEEE (IEEE, 2014)
pp. 6919 -- 6922.
[31] Y. Kobayashi, J. O. J. Okamoto, and M. Fujie, Pro-
ceedings of the 2005 IEEE International Conference on
Robotics and Automation , 1644 (2005).
[32] Y. Kobayashi, A. Onishi, T. Hoshi, K. Kawamura,
M. Hashizume, and M. G. Fujie, International Journal of
Computer Assisted Radiology and Surgery 4, 53 (2009).
[33] Y. Kobayashi, A. Kato, H. Watanabe, T. Hoshi,
K. Kawamura, and M. G. Fujie, Journal of Biomechan-
ical Science and Engineering 7, 177 (2012).
[36] M. Tsukune, Y. Kobayashi, T. Miyashita, and G. M.
Fujie, International Journal of Computer Assisted Radi-
ology and Surgery 10, 593 (2014).
[37] Y. Kobayashi, M. Suzuki, A. Kato, M. Hatano, K. Kon-
ishi, M. Hashizume, and M. G. Fujie, IEEE Transactions
on Robotics 28, 710 (2012).
[38] S. Das, Functional fractional calculus (Springer Science
& Business Media, 2011).
[39] A. W. Lohmann, D. Mendlovic, and Z. Zalevsky, Optics
letters 21, 281 (1996).
[40] H. Ono, Journal of the Physical Society of Japan 39, 1082
(1975).
[41] T. B. Benjamin, Journal of Fluid Mechanics 29, 559
(1967).
[42] J. F. Kelly and R. J. McGough, The Journal of the
Acoustical Society of America 126, 2072 (2009).
[43] T. Hoshi, Y. Kobayashi,
and M. G. Fujie, Proceed-
ings of the 2nd Biennial IEEE/RAS-EMBS International
Conference on Biomedical Robotics and Biomechatron-
ics, BioRob 2008 , 730 (2008).
|
1112.0876 | 1 | 1112 | 2011-12-05T10:10:47 | Spontaneous helicity of a polymer with side-loops confined to a cylinder | [
"physics.bio-ph",
"cond-mat.soft"
] | Inspired by recent experiments on the spatial organization of bacterial chromosomes, we consider a type of "bottle brush" polymer consisting of a flexible backbone chain, to which flexible side loops are connected. We show that such a model with an open linear backbone spontaneously adopts a helical structure with a well-defined pitch when confined to small cylindrical volume. This helicity persists over a range of sizes and aspect-ratios of the cylinder, provided the packing fraction of the chain is suitably large. We analyze this results in terms of the interplay between the effective stiffness and actual intra-chain packing effects caused by the side-loops in response to the confinement. For the case of a circular backbone, mimicking e.g. the E. coli chromosome, the polymer adopts a linearized configuration of two parallel helices connected at the cylinder poles. | physics.bio-ph | physics |
Spontaneous helicity of a polymer with side-loops confined to a cylinder
FOM Institute AMOLF, Science Park 104, 1098XG Amsterdam, The Netherlands
Debasish Chaudhuri∗ and Bela M. Mulder†
(Dated: May 22, 2018)
Inspired by recent experiments on the spatial organization of bacterial chromosomes, we consider
a type of "bottle brush" polymer consisting of a flexible backbone chain, to which flexible side loops
are connected. We show that such a model with an open linear backbone spontaneously adopts a
helical structure with a well-defined pitch when confined to small cylindrical volume. This helicity
persists over a range of sizes and aspect-ratios of the cylinder, provided the packing fraction of the
chain is suitably large. We analyze this results in terms of the interplay between the effective stiffness
and actual intra-chain packing effects caused by the side-loops in response to the confinement. For
the case of a circular backbone, mimicking e.g. the E. coli chromosome, the polymer adopts a
linearized configuration of two parallel helices connected at the cylinder poles.
PACS numbers: 82.35.Pq, 87.16.Sr, 87.15.ap, 89.75.Fb
That confinement can dramatically influence the prop-
erties of non-ideal polymers has already been appreciated
theoretically for a long time. Scaling arguments suggest
that when a confining dimension D becomes smaller than
the natural size of a linear polymer, typically given by its
radius of gyration R, the polymer, rather than behaving
like a single coherent 'blob' of size R, effectively becomes
a string of 'blobs' of size D [1], signalling a cross-over
to lower-dimensional behavior. The experimental explo-
ration of this regime, however, is difficult using synthetic
polymers, whose typical radii of gyration are in the nm
regime and moreover tend to be highly polydisperse in
length [2]. Fortunately nature provides us with an im-
FIG. 1:
(Color online) Helical equilibrium structure for a
backbone chain of length lb = 200σ [black (blue) thick line] to
which side-loops of length ls = 40σ are attached at each back-
bone monomer. The polymer is confined within a cylinder of
length L = 50.75σ and diameter D = 29.5σ. The side-loop
monomers are shown as transparent gray (green) beads.
portant class of model polymers in the form of DNA. To
start with DNA is produced with almost perfect length
control. A well-studied example in vitro is purified λ-
phage DNA which contains exactly 48502 basepairs and
has a length of 16.5 µm and a radius of gyration of 2.1
µm. Using biochemical tools this length can be scaled
to multiples of this unit. Moreover, DNA can readily be
fluorescently labeled. These two properties of DNA were
exploited in a number of recent experiments aimed at
studying the behavior of polymers confined to nanoflu-
idic channels [3–5].
More importantly, in vivo DNA in the form of chro-
mosomes is almost invariably strongly confined. The
paradigmatic case is the 1.5 mm long circular chromo-
some of the bacterium E. coli, which is confined to a
cylindrical volume of diameter ∼ 0.8µm and length vary-
ing between 2 − 4µm. Over the past few years there has
been increasing interest in understanding the details of
the chromosomal configuration in E. coli and its impli-
cations for the, as yet far from fully understood, mech-
anism of sister-chromosome segregation prior to division
[6–8]. Strikingly, recent microscopic observations have
shown that bacterial chromosomes can exhibit a helical
spatial density distribution within the cellular confine-
ment, with a pitch-length in the order of a fraction of
the cell length [9, 10].
Here our aim is to explore to what extent the physics of
highly confined polymers by itself provides clues to such
novel global configurations. As a first step in that direc-
tion it is necessary to take into account that the structure
of a typical bacterial chromosome is not simply that of
a closed linear polymer chain. The combined effects of
supercoiling, due to a globally maintained under-twist
of the DNA, the action of a number of chromosome re-
modelling proteins, or even electrostatic "zippering" can
cause loops in the DNA [11], indicating that a more com-
plex structural picture is needed. Although in reality
the dynamics of the formation and the statistics of such
loops are likely to be highly complex, almost certainly
involving polydispersity in loop sizes and topological en-
tanglements, we chose to focus on arguably the simplest
model of a polymer "dressed" by a cloud of loops, i.e. a
backbone chain to which side-loops of equal size are at-
tached at a regular spacing. The same model has recently
also been discussed by Reiss et al. [12], who have stud-
ied its behavior in the absence of confinement effects.
This type of polymer model is similar to the so-called
"bottle-brush" polymers, extensively studied by Binder
and co-workers [13–15]. The latter work has shown that
such polymers develop a local resistance to bending due
to the entropic repulsion between the side chains. This
effective stiffening, combined with intra-chain packing ef-
fects within the cylindrical confinement, leads, as we will
show in the following, to the spontaneous formation of
helical configurations.
Our model chromosomes are of the bead-spring type,
with consecutive beads attached to each other by a
harmonic spring Vb = (A/2)(di − σui)2 where di =
ri+1 − ri, ri is the position of i-th bead, σ the equi-
librium bond-length and ui = di/di is the local tan-
gent vector to the chain. Non-bonded beads repel each-
other through the Weeks-Chandler-Andersen (WCA) po-
tential [16] VW CA = 4ǫ (cid:2)(σ/rij )12 − (σ/rij )6 + 1/4(cid:3) if the
inter-monomer separation rij < 21/6σ else VW CA = 0,
where ǫ and σ set the energy and length scale of the
system respectively. We use A = 100ǫ. The interac-
tion of all beads with the confining walls are modelled
through Vwall = 2πǫ[(2/5)(σ/riw)10 − (σ/riw)4 + 3/5] if
the distance of the i-th monomer from a wall riw < σ
and Vwall = 0 otherwise. We simulate this system em-
ploying a velocity-Verlet molecular dynamics scheme in
presence of a Langevin thermostat fixing the temperature
at kBT = 1 as implemented by the ESPResSo package
[17].
We first consider a polymer composed of a linear back-
bone chain of length lb = 200σ to which side-loops of
length ls = 40σ are attached at every backbone monomer
of the main chain. This polymer is confined to a cylinder
of length L = 50.75σ and diameter D = 29.5σ, yielding
a monomer packing fraction of η = 23.8%. In Fig. 1 we
show a typical equilibrium configuration of this polymer,
which evidently displays a marked helical ordering of the
backbone chain. The degree of helical ordering can be
quantified by considering the tangent-tangent correlation
function hu(s) · u(0)i, where the positional coordinate is
given by s = iσ with i = 0, 1, . . . , 200. The Fourier trans-
form of this quantity yields a structure function S(q) with
a peak at a dimensionless wavenumber qmax = lb/λmax,
where λmax is the pitch of the helix measured along the
backbone chain. The height of the structure function at
its maximum is a relative measure of the degree of he-
licity, whilst the width of the peak is indicative of the
statistical dispersion of the structure.
Fig. 2 shows the correlation function and the corre-
sponding structure function for this polymer, as we vary
the diameter of the confining cylinder. This shows that
1.2
1
0.8
)
q
(
S
0.6
0.4
0.2
0
0
2
D=14.68σ
D=19.85σ
D=29.5σ
D=58.93σ
Gauss-core
1
0.5
0
-0.5
>
)
0
(
u
.
)
s
(
u
<
-1
0
50
100
s/σ
150
200
4
8
q
12
16
FIG. 2: (Color online) The tangent-tangent correlation (inset)
and its Fourier transform for the polymer shown in Fig. 1.
The correlation function is oscillatory, with the periodicity
captured by the peak in the structure factor at qp = 4 for
D/σ = 14.68, 19.85, 29.5, and qp = 2 for D/σ = 58.93. Also
shown are the corresponding results for a main-chain polymer
with an effective Gaussian-core mimicking the effect of inter-
side-chain repulsion (cf. Fig. 4(b)).
the helical pitch is relatively robust against changes in
the diameter, although a slight decrease in the amplitude
is apparent as we increase the diameter. Only for the
largest diameter, when both the degree of confinement
as well as the overall packing fraction are significantly
decreased, do we see a preference for a more longitudinal
packing of the main chain.
We now argue that the helical arrangement is stabilized
by two effects. The first is the effective stiffness induced
in the backbone by the presence of side loops. To quan-
tify this intrinsic effect in the absence of confinement we
study a free polymer of length lb = 500, with side-loops
of the same length as before, ls = 40σ, grafted at each
backbone monomer. Although for a very long backbone
(lb ≫ ls) one expects the exponent b, which governs the
mean-square separation between monomers at a distance
s along the backbone through the scaling hr(s)2i ∼ s2b,
to reproduce the value b = 3/5 of a simple self-avoiding
polymer, we find at the shorter length-scales relevant to
our confined polymer a much higher value of b = 0.97
(Fig.3, inset). At the same time, the tangent-tangent
correlation hu(s).u(0)i (Fig.3) shows an extremely weak
power-law decay s−α with α = 0.06, which satisfies the
generic scaling rule α = 2 − 2b. The small value of the
exponent α ≪ 1 suggests that at these relatively short
length scales the backbone is characterized by a signifi-
cant effective stiffness, even close to that of a of a rigid
rod for which b = 1. The algebraic nature of the corre-
lation decay, however, precludes a simple interpretation
in terms of an intrinsic length, in contrast to the persis-
tence length of a chain with intrinsic resistance to bend-
data
s-α
(a)
3
(b)
1
>
)
0
(
u
.
)
s
(
u
<
1000
)
s
(
R
100
10
1
1
0.1
1
data
s b
100
1000
s/σ
10
s/σ
10
100
1000
FIG. 3: (Color online) The tangent-tangent correlation shows
a weak power-law decay s−α with α = 0.06. Inset: power-law
growth of distance between two segments R(s) = phr(s)2i ∼
sb with b = 0.97.
ing. Nevertheless, it is intuitively clear that there is a
free-energy cost associated with local deformations of the
backbone caused by the changes induced in the side-loop
packing. For comparison, we note that the end-to-end
distance measured for our backbone with lb = 500, which
we determined to be R ≈ 413 σ, would be reproduced by
a worm-like chain with persistence length P ≈ 391 σ [18]
comparable to the contour length.
However, backbone stiffness alone is not sufficient to
explain the emergent helicity. Although earlier work has
shown that a persistent chain confined to the surface of a
cylinder can adopt helical confirmation [19], a persistent
chain confined to the volume of a cylinder does not. We
confirm this by simulating a worm-like-chain (WLC) of
length lb = 200 σ with a persistence length of P = 2lb,
close to the effective value determined above for the free
polymer. A typical equilibrium structure of the confined
WLC is shown in Fig. 4, which clearly displays the ten-
dency of such a chain to align itself with the long-axis
of the cylinder, without much internal structure develop-
ing, also supported by an analysis of the tangent-tangent
correlations (see supplementary material).
This negative result for the persistent chain, points
to the importance of the packing of the side-loops in
stabilizing the helical structure. One could naively ex-
pect that the interactions between the side loops cause
an additional effective soft repulsion between the back-
bone monomers with a range determined by the radius of
gyration of the side-loops. This repulsion has a dual role;
as we have shown above, it stiffens the backbone at short
length scales, but at long length scales it acts to keep
distant parts of the backbone apart, effectively thicken-
ing the backbone to a soft 'tube'. To assess the validity
of this latter idea we consider a chain whose monomers,
apart from interacting through the short range WCA po-
FIG. 4: (Color online) (a) Self-avoiding WLC and (b) self-
avoiding Gaussian-core polymer within a cylinder. The chain
has contour length lb = 200σ, and for the WLC polymer the
persistence length P = 2lb. The confining cylinder is of length
L = 50.75σ and diameter D = 29.5σ for both (a) and (b).
tential, also interact with an effective Gaussian core in-
teraction
βVgc(r) = a exp[−(r/w)2]
(1)
intended to mimic, as simply as possible, the soft repul-
sion between the side-loops. Bolhuis et al. showed that,
in free space, the effective interaction between two linear
polymers of the same length can be approximated by the
above expression with a ∼ 2 and w = Rg, the radius of
gyration of the individual polymers [20]. While it is well
known that the free energy cost of overlap between two
polymers in free space is independent of polymer size [21–
23], the blob picture of polymers suggests that their over-
lap free energy will become a linear function of polymer
length [7] when the chains are strongly confined. We
therefore take the interaction strength a = 40 to mimic
the densely-packed side-loops of size ls = 40σ. From our
simulations of side-loop coupled back-bone, we found the
mean radius of gyration Rg = 2.95 σ of the side-loops of
length ls = 40σ. However, due to the repulsive inter-
actions with the backbone monomers the center of mass
of the loops is offset from the backbone. We therefore
extracted the distribution of the center of mass of the
side-loops with respect to their attachment point on the
backbone in the plane perpendicular to the local tangent
direction along the chain (see Figure 6 in the Supplemen-
tary Material). This distribution has a sharp maximum
at a distance of approximately 4σ from the backbone
chain. This result also shows that indeed the density dis-
tribution of the side-loop monomers is structured in a
manner consistent with the 'thickened tube' picture al-
luded to above. We therefore took the size parameter
characterizing the range of the interaction between the
side-loops in the effective potential centered on the back-
bone monomers Vgc to be w ≃ Rg + 4σ = 6.95σ. This ef-
fective interaction between the monomers was then added
to the WCA potential, governing the non-bonded repul-
4
fects caused by side-loops in polymers leads to novel
helical equilibrium configurations of confined polymers.
These structures are strikingly similar to ones recently
observed in bacterial nucleoids. To what extent the phys-
ical effects discussed here are sufficient to explain all
the details of the large scale chromosome organisation
in real bacteria is a question which clearly requires fur-
ther research. At the very least, however, our results
once again indicate that the ubiquitous aspecific interac-
tions between the segments of long biopolymers like DNA
can by themselves lead to significant spatial structuring,
as has previously also been observed in the context of
chromosome organisation in the nuclei of plants [24] and
humans [25]. This argues for a more prominent place for
polymer physics in the research into the structure and
function of chromosomes. From a purely physical point
of view our work points to novel possibilities for "sculpt-
ing" the configurations of confined polymers by judicial
choices of polymer topologies.
We gratefully acknowledge discussions with Nancy
Kleckner and Mara Prentiss (Molecular and Cellular Bi-
ology, Harvard) in the initial stages of this project. This
work is part of the research program of the "Sticht-
ing voor Fundamenteel Onderzoek der Materie (FOM)",
which is financially supported by the "Nederlandse Or-
ganisatie voor Wetenschappelijk Onderzoek (NWO)".
The work of DC was supported by FOM-programme
Nr. 103 "DNA in action: Physics of the genome".
∗ Electronic address: [email protected]
† Electronic address: [email protected]
[1] P.-G. de Gennes, Scaling concepts in polymer physics
(Cornell University Press, 1979).
[2] J. Brandrup, E. H. Immergut, and E. A. Grulke, eds.,
Polymer Handbook, 4th ed. (Wiley-Interscience, 1999).
[3] D. J. Bonthuis, C. Meyer, D. Stein, and C. Dekker, Phys.
Rev. Lett., 101, 108303 (2008).
[4] J. Tang, S. L. Levy, D. W. Trahan, J. J. Jones, H. G.
Craighead, and P. S. Doyle, Macromolecules, 43, 7368
(2010).
[5] H. Uemura, M. Ichikawa, and Y. Kimura, Phys. Rev. E,
81, 051801 (2010).
[6] D. Bates and N. Kleckner, Cell, 121, 899 (2005).
[7] S. Jun and B. Mulder, Proceedings of the National
Academy of Sciences of the United States of America,
103, 12388 (2006).
[8] P. Wiggins, K. Cheveralls, J. Martin, R. Lintner, and
J. Kondev, Proceedings of the National Academy of Sci-
ences, 107, 4991 (2010).
[9] I. A. Berlatzky, A. Rouvinski,
and S. Ben-Yehuda,
Proceedings of the National Academy of Science, 1051,
14136 (2008).
[10] C. Butan, L. M. Hartnell, A. K. Fenton, D. Bliss, R. E.
Sockett, S. Subramaniam, and J. L. S. Milne, Journal of
bacteriology, 193, 1341 (2011).
[11] S. B. Zimmerman, Journal of Structural Biology, 156,
FIG. 5: (Color) Equilibrium structure for a circular backbone
chain of length lb = 400σ, with side-loops of length ls =
20σ attached to each backbone monomer. The polymer is
confined within a cylinder of length L = 50.4σ and diameter
D = 33.5σ. The splitting of the backbone into two parallel
linearized branches is highlighted by using two color-codes,
orange and blue, for the branches. The side-loop monomers
are shown as transparent green beads.
sive interactions, for a linear chain of length lb = 200σ
confined within a cylinder of diameter D = 29.5σ and
length L = 50.75σ. As Fig.4b shows, this effective poten-
tial reproduces the helical equilibrium structures of the
polymer remarkably well. The structure factor displays
a maximum at the same helical pitch value lb/4 = 50σ as
the original simulations and reproduces the oscillations
of the tangent-tangent correlation function (Fig. 2) with
only a slight global phase shift. The amplitude of the os-
cillations (and hence that of the structure factor) is some-
what larger for the effective potential, but this is prob-
ably due to an overestimate of the interaction strength
a.
Finally we address what happens if the backbone is a
ring-like polymer, appropriate to modeling, e.g., the cir-
cular chromosome of E. coli. To that end we simulated
a polymer with a circular backbone with lb = 400σ and
side-loops of length ls = 20σ, trapped within a cylinder
of length L = 50.4σ and diameter D = 33.5σ. The pack-
ing fraction of the monomers is η = 18.9%, comparable to
the one in the simulation of the linear backbone polymer.
Fig. 5 shows that in this case the backbone loop is now
organized into two parallel helices running along the long
axis of the cylinder. As is evident from the snapshot, and
corroborated by the analysis of the tangent-tangent cor-
relations (Fig. 8 in Supplementary Material), the degree
of helicity is reduced as compared to the linear backbone
case due to the smaller side-loop length, but nevertheless
remains significant.
In conclusion, we have shown that the interplay be-
tween the effective stiffness and intra-chain packing ef-
5
255 (2006).
[12] P. Reiss, M. Fritsche, and D. W. Heermann, Phys. Rev.
E, 84, 051910 (2011).
[13] R. Wang, P. Virnau, and K. Binder, Macromolecular
Theory and Simulations, 19, 258 (2010).
[14] H.-P. Hsu, W. Paul, and K. Binder, Macromolecules, 43,
3094 (2010).
[15] P. E. Theodorakis, H.-P. Hsu, W. Paul, and K. Binder,
The Journal of chemical physics, 135, 164903 (2011).
[16] J. D. Weeks, D. Chandler, and H. C. Andersen, J. Chem.
Phys., 54, 5237 (1971).
[17] H.-J. Limbach, A. Arnold, B. A. Mann, and C. Holm,
Comput. Phys. Commun., 174, 704 (2006).
[18] J. Hermans and R. Ullman, Physica, 18, 951 (1952).
[19] I. Kusner and S. Srebnik, Chemical Physics Letters, 430,
84 (2006).
[20] P. G. Bolhuis, A. A. Louis, J. P. Hansen, E. J. Meijer,
Journal of Chemical Physics, 114, 4296 (2001).
[21] J. des Cloizeaux, Journal de Physique, 36, 281 (1975).
[22] M. Daoud, J. P. Cotton, B. Farnoux, G. Jannink,
G. Sarma, H. Benoit, C. Duplessix, C. Picot, and P. G.
de Gennes, Macromolecules, 8, 804 (1975).
[23] A. Y. Grosberg, P. G. Khalatur, and A. R. Khokhlov,
Die Makromolekulare Chemie, Rapid Communications,
3, 709 (1982).
[24] S. de Nooijer, J. Wellink, B. Mulder, and T. Bisseling,
Nucl. Acids Res., 37, 3558 (2009).
[25] P. R. Cook and D. Marenduzzo, The Journal of cell bi-
ology, 186, 825 (2009).
SUPPLEMENTARY: SPONTANEOUS HELICITY
OF A POLYMER WITH SIDE-LOOPS
CONFINED TO A CYLINDER
0.08
0.06
)
r
(
P
0.04
0.02
P(rn)
P(rt)
0.2
0.15
)
q
(
S
0.1
0.05
0
0
6
>
)
0
(
u
.
)
s
(
u
<
1
0.8
0.6
0.4
0.2
0
-0.2
0
50
100
s/σ
150
200
10
20
q
30
40
0
0
2
4
r/σ
6
8
FIG. 6: The distribution of center of mass distance of the side-
loops around the main-chain. rt denotes the component of
the distance along the local tangent of the main chain and rn
denotes the component in a plane perpendicular to the local
tangents. P (rn) has a pronounced maximum near rn = 4σ,
whereas P (rt) has a Gaussian shape with a variance of 8σ2.
FIG. 8: The tangent-tangent correlation and its Fourier trans-
form for a circular main-chain polymer attached with side-
loops – the main-chain has length lb = 200σ and side-loops
of length ls = 40σ – and confined within a cylinder of length
L = 50.4σ and diameter D = 33.5σ. The correlation (inset)
shows clear periodic oscillation, and the periodicity is cap-
tured by the peak in the Fourier transform at qp = 4.
1
0.8
0.6
0.4
0.2
)
q
(
S
WLC
1
0.5
0
-0.5
>
)
0
(
u
.
)
s
(
u
<
0
50
100
s/σ
150
200
0
0
4
8
q
12
16
20
FIG. 7: The tangent-tangent correlation (inset) and its
Fourier transform for a WLC polymer in presence of WCA
repulsion between non-bonded monomers.
We calculate the probability distribution of the center
of mass position of the side-loops around the main chain
polymer (Fig. 6). This is done using our direct simu-
lation of a main-chain polymer attached with side-loops
confined within a cylinder. The main-chain has length
lb = 200σ to each monomer of which is attached side-
loops of length ls = 40σ. The polymer is confined within
a cylinder of length L = 50.75σ and diameter D = 29.5σ.
The center of mass position of the loop r is measured
from the main-chain monomer to which the loop is at-
tached. This position vector can be projected onto the
direction of a local tangent on the main chain t, and
t, where
a normal direction n such that r = rn n + rt
rn = r.n and rt = r.t. We calculate the distribution
P (rt), P (rn) using 1000 equilibrated configurations of
the polymer (Fig. 6).
We find that the center of mass of the side chains
are distributed in a Gaussian manner in the direction
parallel to the local tangent P (rt). Howeever, in the
perpendicular planes, the distribution P (rn) has a pro-
nounced maximum at a distance rn = 4σ. This off-center
location of the center of mass of the side-loops, along
with the radius of gyration Rg = 2.95 σ leads us to use
w ≃ Rg + 4σ = 6.95σ for the effective inter-side-loop
interaction βVgc(r) = a exp[−(r/w)2].
In Fig. 7 we plot the tangent-tangent correlation and
its Fourier transform for a self-avoiding (non-bonded
monomers repel by WCA potential) WLC polymer of
chain-length lb = 200 σ and persistence length P = 2lb,
confined within a cylinder of length L = 50.75 σ and di-
ameter D = 29.75 σ (corresponding to Fig.4(a) of main
text). Note the reduced magnitude and pitch of the helic-
ity comapred to that of self-avoiding polymer in presence
of the Gaussian-core repulsion as shown in Fig.1 of main
text.
In Fig. 8 we show the tangent-tangent correlation and
its Fourier transform for the case when the backbone is
a ring-like polymer shown in Fig.5 of main text. The
correlation function shows nice periodic oscillation which
is captured by the maximum in the Fourier transform
at qp = 4. This peak corresponds to a helical pitch of
λmax = lb/4 = 50 σ.
7
|
1902.10323 | 1 | 1902 | 2019-02-27T04:12:39 | Quantifying Dissipation in Actomyosin Networks | [
"physics.bio-ph"
] | Quantifying entropy production in various active matter phases will open new avenues for probing self-organization principles in these far-from-equilibrium systems. It has been hypothesized that the dissipation of free energy by active matter systems may be optimized to produce highly dissipative dynamical states, hence, leading to spontaneous emergence of more ordered states. This interesting idea has not been widely tested. In particular, it is not clear whether emergent states of actomyosin networks, which represent a salient example of biological active matter, self-organize following the principle of dissipation optimization. In order to start addressing this question using detailed computational modeling, we rely on the MEDYAN simulation platform, which allows simulating active matter networks from fundamental molecular principles. We have extended the capabilities of MEDYAN to allow quantification of the rates of dissipation resulting from chemical reactions and relaxation of mechanical stresses during simulation trajectories. We validate our approach with a mean-field model that estimates the rates of dissipation from filament treadmilling. Applying this methodology to the self-organization of small disordered actomyosin networks, we find that compact and highly cross-linked networks tend to allow more efficient transduction of chemical free energy into mechanical energy. In these simple systems, we do not observe that spontaneous network reorganizations lead to increases in the total dissipation rate as predicted by the dissipation-driven adaptation hypothesis mentioned above. However, whether such a principle operates in more general, more complex cytoskeletal networks remains to be investigated. | physics.bio-ph | physics | Quantifying Dissipation in Actomyosin Networks
Carlos Floyd1, Garegin A. Papoian1,2,3,†, and Christopher Jarzynski1,2,3,4,*
1Biophysics Program, University of Maryland, College Park, MD 20742 USA
2Department of Chemistry and Biochemistry, University of Maryland, College Park,
3Institute for Physical Science and Technology, University of Maryland, College Park,
MD 20742 USA
MD 20742 USA
4Department of Physics, University of Maryland, College Park, MD 20742 USA
†email: [email protected]
*email: [email protected]
February 28, 2019
Abstract
Quantifying entropy production in various active matter phases will open new avenues
for probing self-organization principles in these far-from-equilibrium systems. It has been
hypothesized that the dissipation of free energy by active matter systems may be optimized
to produce highly dissipative dynamical states, hence, leading to spontaneous emergence of
more ordered states. This interesting idea has not been widely tested. In particular, it is not
clear whether emergent states of actomyosin networks, which represent a salient example
of biological active matter, self-organize following the principle of dissipation optimization.
In order to start addressing this question using detailed computational modeling, we rely
on the MEDYAN simulation platform, which allows simulating active matter networks from
fundamental molecular principles. We have extended the capabilities of MEDYAN to allow
quantification of the rates of dissipation resulting from chemical reactions and relaxation
of mechanical stresses during simulation trajectories. This is done by computing precise
changes in Gibbs free energy accompanying chemical reactions using a novel formula, and
through detailed calculations of instantaneous values of the system's mechanical energy.
We validate our approach with a mean-field model that estimates the rates of dissipation
from filament treadmilling. Applying this methodology to the self-organization of small
disordered actomyosin networks, we find that compact and highly cross-linked networks
tend to allow more efficient transduction of chemical free energy into mechanical energy.
In these simple systems, we do not observe that spontaneous network reorganizations lead
to increases in the total dissipation rate as predicted by the dissipation-driven adaptation
hypothesis mentioned above. However, whether such a principle operates in more general,
more complex cytoskeletal networks remains to be investigated.
9
1
0
2
b
e
F
7
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
3
2
3
0
1
.
2
0
9
1
:
v
i
X
r
a
1
1
Introduction
The actin-based cytoskeleton is a dynamic supramolecular structure that, by sustaining and
releasing mechanical stress in response to various physiological cues, mediates the exertion of
force by cells both on their environments and within their bodies [1, 2]. These cytoskeletal
structures are composed of long (on the order of 1 µm in vivo [3]) actin polymers which are
interconnected by various cross-linkers, as well as by myosin motor filaments, resulting in a three-
dimensional network-like organization referred to as an "actomyosin network" [4, 5].1 Part of
the intricacy of actomyosin network dynamics is due to the mechanosensitive kinetic reaction
rates controlling cross-linker and myosin filament unbinding as well as myosin filament walking:
at high tension, cross-linkers will unbind more quickly (slip-bond) whereas motors will unbind
and walk less quickly (catch-bond and stalling). [6, 7, 8]. These reactions control the actomyosin
network connectivity, which in turn determines the ability of the network to globally distribute
stress [9]. Thus the mechanosensitive feedback introduces nonlinear coupling between the stress
sustained by an actomyosin network and the network's ability to reorganize in response to that
stress. In order to be responsive to physiological cues, the dynamics of these systems occur far
from thermodynamic equilibrium; the hydrolysis of an out-of-equilibrium concentration of ATP
molecules fuels a) the stress-generating activity of the myosin motor filaments, and b) filament
treadmilling [10, 11, 12, 13, 14]. Filament treadmilling is a steady-state situation in which the
polymerization at the plus end of the filament is compensated by the depolymerization at the
minus end, resulting in the filament moving forward without its length changing. As a result of
these these local free energy-consuming processes, actomyosin networks constitute an interesting
and biologically important example of soft active matter. Active matter is composed of agents
that individually transduce free energy from some external source, in this case the chemical
potential energy of many ATP molecules [15, 16, 17]. Dissipation in these systems results when
the free energy consumed ∆G is greater than the quantity of work W done by the system on
its environment, with the remainder ∆G − W serving to increase the total entropy.
The viewpoint of actomyosin networks as active matter systems has been fruitfully
adopted in recent theoretical and experimental studies, yet a lack of ability to quantify the
rates of free energy transduction by these systems has hindered development of some of these
lines of study. The emergence of distinctive dynamical states (for instance pulsing actin waves or
vortices) during the self-organization of actomyosin systems has been documented in several in
vitro experiments [18, 19, 20]. These emergent patterns depend sensitively on the concentrations
of myosin filaments and cross-linkers: myosin filament concentration controls the level of active
stress generation, and cross-linker concentration controls the degree of mechanical coupling
of actin filaments, which has been described using the language of percolation theory [21, 9].
While these emergent dynamic patterns have been characterized in detail, a general mechanism
explaining why these patterns emerge under given conditions has not yet been proven.
It
might be expected, given that these systems operate away from thermodynamic equilibrium,
that the quantity of free energy dissipated during a system's evolution is optimized, similar
to the principle of minimum entropy production in the near-equilibrium theory of irreversible
thermodynamics [22]. However, this minimum entropy production principle breaks down in the
far-from-equilibrium, nonlinear-response regime, where many active matter systems including
actomyosin networks operate [23].
It has recently been proposed that another optimization
principle applies arbitrarily far from equilibrium. This idea, referred to as dissipation-driven
adaptation, suggests that, in general, a coarse-grained trajectory of some nonequilibrium system
will be more likely than all alternative trajectories if the amount of free energy absorbed and
dissipated along that trajectory is maximal [24, 25, 26]. This organizing principle has been borne
1In our terminology, we will distinguish between "cross-linking proteins," which will include both active (e.g.
myosin filaments) and passive (e.g. α-actinin and fascin) proteins that bind to adjacent actin filaments, and
"cross-linkers," which refer exclusively to passive cross-linking proteins.
1
out in model systems [27, 28], yet has also been shown to have certain counter examples [29]; it
remains actively debated. In the case of actomyosin systems, it has not yet been tested because
of the difficulty in measuring dissipated free energy using most experimental approaches. In
this paper we take the first steps toward such a test, by developing a simulation methodology
allowing the quantification of dissipated free energy during the self-organization of actomyosin
networks.
In addition to being of interest in the field of active matter systems, dissipation in
actomyosin networks has also been an important factor in recent experimental studies of cell
mechanics. Rheological properties of actomyosin networks largely determine rheological prop-
erties of the whole cell, and it has been discovered that the dissipative component of the cell's
viscoelastic response to mechanical oscillations (called the loss modulus) is partly attributable
to the ATPase activity of myosin motor filaments in actomyosin cytoskeleton [30, 31, 32]. In the
context of cell mechanics, myosin filaments have several roles: they produce mechanical stress by
pulling on actin filaments, they dissipate mechanical stress by disassembling actin filaments and
higher order stress-sustaining filament structures, and they consume chemical energy through
the hydrolysis of ATP molecules. It has been proposed that a lack of detailed understanding of
the effects of these processes, and of dissipation in actomyosin systems more generally, under-
lies inconsistent, widely variable traction force microscopy measurements of cell migration [33].
Progress along this line is hindered by an absence of methods to study dissipation in actomyosin
networks directly and at sufficiently high spatio-temporal resolution.
To address these needs, we introduce a computational approach to measure dissipa-
tion during simulations of actomyosin network self-organization using the simulation platform
MEDYAN (Mechanochemical Dynamics of Active Networks) [34, 35]. MEDYAN simulations
marry stochastic reaction-diffusion chemistry algorithms with detailed mechanical models of
actin filaments, cross-linkers, myosin motor filaments, and other associated proteins, and it also
accounts for mechanosensitive reaction rates. This combination of simulation features makes
this software uniquely capable of probing the complexity of actomyosin network dynamics. For
instance, past studies utilizing MEDYAN have investigated the dependence of network collapse
on myosin filament and cross-linker concentrations, as well as the origin of local contractility
in actomyosin networks [34, 36]. We refer the reader to the paper describing MEDYAN for a
detailed discussion of the various aspects of the simulation platform [34], while here we describe
an extension of that platform that allows for calculation of the energetics of the chemical and
mechanical events occurring during simulation. We utilize these new capabilities to characterize
the dissipation resulting from filament treadmilling, for which we further introduce a mean-field
model, as well as from myosin filament walking. We study both the time-dependence and the
distributions of dissipation rates as concentrations of cross-linkers and myosin filaments are
varied, observing that transduction of chemical energy to stored mechanical energy is more
efficient at denser network organizations. For these simulations, we first explore systems with
"plain" myosin filaments and cross-linkers that are not mechanosensitive, in order to simplify
the overall dynamics. We then introduce their mechanochemical coupling to understand its
effect on the observed trends. We end by discussing how this new methodology can provide a
valuable technique to advance the studies of actomyosin networks mentioned above.
2 Methods
2.1 Measuring Dissipation in MEDYAN
We first give a brief overview of the MEDYAN simulation platform. MEDYAN employs a
stochastic chemical evolution algorithm in conjunction with mechanical representations of poly-
mers and cross-linking proteins to simulate the dynamics of networks with active components,
including but not limited to actomyosin networks. The simulation space comprises a grid of
2
reaction-diffusion compartments, inside which chemical species (e.g. unpolymerized subunits
or cross-linking proteins) are assumed to be homogeneously distributed without specified lo-
cations, and which participate in reactions (e.g. (de)polymerization or (un)binding) according
to mass-action kinetics; the species can additionally jump between compartments in diffusion
events. When an unpolymerized subunit polymerizes to or nucleates a filament, it becomes
part of the mechanical subsystem, gaining location coordinates in the simulation volume and
becoming subject to mechanical potentials depending on its interaction with other mechanical
elements. Through chemical reactions such as myosin filament binding and walking, the me-
chanical energy of the system changes, and the new net forces are then periodically relaxed
in a mechanical equilibration phase, using conjugate gradient minimization. We fill in salient
details of the above overview as they become relevant below. A user provides input data in-
cluding system size and simulation length, mechanical parameters (e.g. stretching and bending
constants and excluded volume cutoff distances), size of the polymer subunits, energy minimiza-
tion algorithm parameters, chemical simulation algorithm parameters, choices for the modeling
of force-sensitive reaction rates, initial conditions of the filaments (either specified or randomly
generated), a list of reacting species and their associated parameters, a list of reactions involving
those species and their associated parameters, and a list of desired output information. The
output of a simulation is a set of trajectory files containing information at each time point,
which can include positions of the mechanical network elements, tensions on the elements, and
copy numbers of the chemical species, among other things. In Supplementary Information we
discuss parameterization of the simulations analyzed in this paper. MEDYAN is extensible in
that it is possible to implement new types of outputs, depending on the experimental needs; in
this paper we describe a novel output that reports the changes in the Gibbs free energy of the
system.
As a MEDYAN simulation progresses, the Gibbs free energy of the system continually
changes due to occurrences of chemical reactions and structural rearrangements of the polymer
network. These processes are driven by an out-of-equilibrium concentration of ATP which
fuels filament treadmilling and myosin filament walking. Dissipation measurement in MEDYAN
works by calculating running totals of the chemical and mechanical energy changes. The running
totals can then be converted into instantaneous rates by taking the numerical derivative at each
time point using the forward difference quotient. The algorithm for tracking these energy
changes is compatible with the following sequence of consecutive procedures that make up one
iterative cycle of a MEDYAN simulation [34]:
1. Evolve system with stochastic chemical simulation for time tmin.
2. Calculate the changes in the mechanical energy resulting from the reactions in Step 1.
3. Mechanically equilibrate the network based on the new stresses calculated in Step 2.
4. Update the reaction rates of force-sensitive reactions based on the new forces.
Dissipation tracking is done by calculating for each of these four steps a change in free
energy, and then adding these free energy changes to determine the total change in free energy
resulting from each cycle, ∆Gdissipated. Since, at least in this study, the actomyosin network is
not mechanically coupled to any work reservoir external to the simulation volume (i.e. W = 0),
∆Gdissipated is indeed dissipated energy [37]. This methodology could be straightforwardly
extended to account for work exchanged with an external system in future studies, however.
Step 4 in the MEDYAN simulation cycle does not result in a change in free energy, as explained
in Supplementary Information. The notation chosen for these free energy changes, as well as
the direction in which the energy is changing during each procedure, is illustrated in Figure 1.
3
Figure 1: Energy level diagram indicating which changes in free energy are tabulated during the
4 procedures constituting one cycle in MEDYAN simulation. Step 4, mechanochemical update
of reaction rates, does not result in a free energy change. Dotted lines represent energy levels
that are intermediate during the iterative cycle, and solid lines represent the energies at the
beginning and end of one cycle.
From this picture, we have the following relations:
∆Gdissipated = ∆Gchem + ∆Gmech
∆Gdissipated = ∆Gchem, dissipated + ∆Gmech, dissipated
∆Gchem, dissipated = ∆Gchem + ∆Gstress
∆Gmech, dissipated = ∆Gmech − ∆Gstress
(1)
(2)
(3)
(4)
For the depicted relative position of energy levels, the sign convention is such that all values of
∆G except for ∆Gstress will be negative (indicated by the arrow's direction), since G refers to
the free energy of the system, not of its environment, and will therefore be tend to be negative as
the system moves down the free energy landscape. The usual intuition that the total dissipation
is positive can be stated
∆G+
dissipated = −∆Gdissipated > 0,
(5)
where the superscript "+" indicates the positive change in the total entropy.
Equation 3 says that, given some change in the system's chemical potential energy,
∆Gchem, resulting from reactions occurring during Step 1, a portion of that energy is used
to deform the polymer network (e.g. via myosin filament pulling on actin filaments). This
increases the mechanical energy in the network by an amount ∆Gstress. Only the portion of
∆Gchem which has not gone into ∆Gstress has been dissipated as heat. In Step 3 the network is
mechanically equilibrated, resulting in relaxation of net forces (though not of all stresses) and
updating of the network elements' positions. We refer to the decrease in mechanical energy
resulting from this relaxation as ∆Gmech, dissipated.
The calculation of ∆Gstress and ∆Gmech, dissipated is based on a set of mechanical poten-
tials describing interactions between elements of the actomyosin network. Polymers are modeled
as a sequence of thin, unbendable, yet extensible cylinders that are joined at their ends by beads
whose positions define the polymer's configuration. The structural resolution of MEDYAN is
at the level of the cylinders, which in this study are 27 nm long and have effective diameters of
approximately 5 nm, however the diameter is not a parameter of the simulation, being instead
effectively determined by the strength of the excluded volume interaction between cylinders.
Cross-linking proteins (e.g. α-actinin and myosin filaments) are modeled as Hookean springs
connecting these cylinders by attaching to discrete binding sites. Included among the mechan-
ical potentials are various modes of filament deformation, excluded volume interactions, and
4
1231 - chemical simulation2 - compute new stresses3 - mechanical equilibration4 - update reaction rates⎨⎩⎧⎜⎜⎨⎩⎧⎨⎩⎧⎨⎩⎧⎨⎩⎧⎜⎜⎩⎧⎜⎨⎜stretching of cross-linkers and myosin filaments. Mechanical equilibration is accomplished by
constrained minimization of the mechanical energy with respect to the positions of the network
elements. A full description of the mechanical potentials and equilibration protocols is given
in [34]. Determining ∆Gstress and ∆Gmech, dissipated requires evaluating the instantaneous total
mechanical energy of the system at certain points during the iterative simulation cycle and
taking the difference of those values.
The calculation of ∆Gchem and ∆Gchem, dissipated is accomplished by incrementing a
running total of the chemical energy Gchem whenever a reaction stochastically occurs during Step
1, and finding the accumulated change at the end of the protocol. Chemical stochastic simulation
in MEDYAN uses a Gillespie-like reaction-diffusion algorithm over a grid of compartments
which constitutes the simulation volume. Diffusing species are assumed to be homogeneous (i.e.
obey mass-action kinetics) inside the compartments, and can jump between the compartments
leading to concentration gradients at the scale of the compartment length (taken to be roughly
the Kuramoto length of diffusing G-actin, following [35]). The evolving polymer network is
overlaid on this compartment grid, with each piece of a polymer reacting with diffusing species
according to the concentrations in its local compartment. Again, we refer the reader to [34] for
a more detailed description of the chemical dynamics. For the present purpose of measuring
dissipation, we introduced into this simulation protocol a precise formula for the change in
Gibbs free energy corresponding to the occurrence of various reactions as a function of the
instantaneous compartment concentrations. In the Supplementary Information we establish this
formula, while we present its derivation in an accompanying paper [38]. The set of chemical
reactions used to describe actomyosin networks in this study is based on a previous model of
actin polymerization dynamics that explicitly treats hydrolysis states of the nucleotide bound to
each actin subunit [39, 40]. This level of detail allows to quantify the dissipation resulting from
ATP hydrolysis during filament treadmilling. To increase computational efficiency, we neglect
the dynamics of nucleotide hydrolysis states of the tips of the filaments. This has been shown
in previous work to be a valid approximation to the full dynamics which includes the states of
the filament tips [40]. We refer to the resulting set of reactions describing actin polymerization
dynamics as the Constant Tip (CT) model. However unlike in the original CT model, here we
explicitly include G-actin bound to ADP-Pi as a reacting species, for completeness and since the
extra computational strain of doing so is small. The actin subunit species tracked in this model
are distinguished by their polymerization state and by the hydrolysis states of the nucleotide to
which they are bound. We notate a species as G or F , to represent globular (un-polymerized)
or filamentous (polymerized) actin respectively, superscripted by T , P i, or D to represent that
it is bound to ATP, ADP-Pi, or ADP, respectively; thus for instance filamentous actin bound
to ADP-Pi is notated F P i. We also include reactions describing cross-linker (un)binding and
myosin filament (un)binding and walking. We exclude filament nucleation, severing, destruction,
and annealing reactions, thus the number of filaments is constant throughout the simulation
trajectories.
In the Supplementary Information we describe how we compute the change in
Gibbs free energy for each reaction in this set, as well as how we parameterize the simulations.
Lastly, we developed a mean-field model to describe just the dissipation resulting from
reactions in the CT model, i.e. excluding cross-linkers and myosin filaments. We describe the
model and present its results in Supplementary Information. We find, among other things, that
the steady-state dissipation rate from filament treadmilling counter-intuitively does not depend
on the total amount of actin, but only on the number of filaments.
5
3 Results
3.1 Total Dissipation Rates of Disordered Networks Do Not Increase
We studied dissipation rates accompanying the process of myosin-driven network self-organization.
We first excluded in these simulations the force-sensitivity of the reaction rates describing cross-
linker unbinding and myosin filament unbinding and walking. This allowed us to understand a
simplified version of the dynamics (we later discuss the the effect of including mechanochem-
ical feedback). We analyzed the trajectories of the quantities ∆Gdissipated, ∆Gchem, dissipated,
∆Gmech, dissipated, ∆Gchem, and ∆Gmech over a set of simulations with identical initial concen-
trations but different random filament distributions. In Figure 2, we display these trajectories
averaged over 10 runs. We used a simulation volume of 1 µm3, divided into 8 compartments,
and initial conditions of equal amounts (10 µM each) of GT and GD actin in a 0.08 µM pool of
seed filaments containing F T , as well as 0.2 µM myosin and 1.0 µM cross-linkers. Similar tra-
jectories for other conditions are shown in SI Figures 2. Points lying outside 3 median absolute
deviations (MAD) have been excluded for this visualization [41] 3.
Figure 2: Top Left: Combined trajectories of 5 quantities tracked during a MEDYAN simulation,
averaged over 10 separate runs. The color coding is indicated by the remaining panels. Note
the close overlap between ∆G+
In the remaining panels, the individual
trajectories are visualized with their standard deviations at each time point over the 10 runs
visualized as lighter curves above and below the main curve. In the plots for ∆G+
dissipated and
∆G+
mech, the full range is visualized, however this range is cropped in the other plots to aid
visibility.
dissipated and ∆G+
chem.
2Due to file size constraints on preprint upload, this supplementary figure has been excluded.
It will be
available in the print version coming soon in Interface Focus.
3For certain statistical aggregations in this paper we use the MAD since it is the estimate of scale most robust
against outliers, having a breakdown point of 50%. However, we use the unscaled version of the MAD and, as a
result, do not claim that this statistic is a consistent estimator of the standard deviation [42]. The distributions
of most of the quantities of interest are too pathological to allow for a straightforward choice of scaling factor,
so we simply us the unscaled version, defined as MAD = medianixi − medianjxj, where xi are elements in the
data set.
6
mech, dissipated to ∆G+
mech, dissipated relative to ∆G+
dissipated tends to increase relative to that of ∆G+
For each tracked quantity, there is an initial transient phase (lasting a few tens of
seconds) followed by fluctuations around a roughly steady value; we do not observe in this
set of simulations any slow approach to a significantly different total dissipation rate, which
might correspond to large reorganizations of the network. However for high concentrations of
cross-linkers (CCL = 5.0 µM ) and myosin (CM = 0.4 µM ), we observe that the contribution of
∆G+
chem, dissipated (SI Figure
2). It is known that network percolation and collapse occurs only under certain conditions of
myosin and cross-linker concentrations [34, 21], and it is under these conditions which result in
collapse that we observe an increase in ∆G+
chem, dissipated (SI Figures
2). This indicates that more mechanical stress is being created by myosin filament walking as
the network collapses and becomes more densely cross-linked. The transient phase corresponds
to the initial polymerization of the seed filaments followed by the initial mechanical coupling of
filaments by cross-linkers and myosin filaments. Following the transient phase, the networks in
this study are generally disordered (Figure 3). The dissipation rate corresponding to the initial
polymerization of the seed filaments is much larger than the chemical dissipation resulting from
myosin filament activity. However, following the transient phase, after which the treadmilling
dissipation has reached its steady-state (SI Figure 9), the contribution from myosin filament
activity outweighs the dissipation resulting from filament treadmilling. Tracking the instan-
taneous rate of change in ∆G+
chem resulting from each reaction separately, we observe myosin
filament walking to contribute the majority to ∆G+
chem after the transient phase, however the
amount depends on CCL and CM . The integrated contributions of each reaction to ∆G+
chem
for different conditions are shown in Figure 4 and SI Figures 2. Interestingly, diffusion con-
tributes an appreciable fraction to ∆G+
chem; plus and minus ends of actin filaments tend to
localize together (Figure 3) and through treadmilling deplete the concentrations of GT and GD
in certain reaction compartments relative to others, leading to significant diffusion gradients
on the scale of the compartment length. This suggests that the establishment of concentration
gradients is an important driving force in actomyosin self-organization, as has been noted in
other work [43, 44]. In the initial transient phase, the mechanical energy changes appreciably
as the filaments grow and are initially coupled to each other. Following this, however, the rate
of ∆G+
mech is on average near zero. This indicates that, despite the process of mechanical stress
generation through myosin filament activity and treadmilling, the resulting stress is dissipated
through fast relaxation such that, on a slower timescale, the mechanical energy of the system
does not change in a significant, persistent way.
7
Figure 3: Snapshot of a percolated actomyosin network in MEDYAN under conditions CCL =
5.0 µM and CM = 0.4 µM . Actin filaments are drawn in red, cross-linkers are drawn in green,
myosin filaments are drawn in blue, and the plus ends of filaments are drawn as black spheres.
8
Figure 4: Integrated contributions of each reaction in MEDYAN to the total ∆Gchem along a
simulation trajectory with CCL = 1.0 µM and CM = 0.2 µM .
Aggregating each of the 2000 s of the 10 trajectories into a collective data set for
each condition of CM and CCL, we next analyzed the distribution of the instantaneous rates of
∆G+
chem. In Figure 5
we plot histograms of these 6 quantities for the conditions CCL = 1.0 µM , CCL = 0.2 µM . In
SI Figures 2 we show the same plots for other conditions.
mech, dissipated, ∆G+
chem, dissipated, ∆G+
dissipated, ∆G+
stress, ∆G+
mech, and ∆G+
9
Figure 5: Histograms and fitted probability distribution functions for 6 tracked quantities. For
each histogram, the full trajectory for each of 10 runs is combined into a single data set. A
log-normal distribution was used to fit the histograms of ∆G+
chem, a generalized
normal distribution was used to fit ∆G+
mech, and the rest were fit with gamma distributions. All
distributions are fit using the SciPy package to determine shape, scale, and location parameters
[45]. Quantities were made positive or negative in order to produce the best fits.
dissipated and ∆G+
Each distribution contained heavy tails which, for the purpose of fitting, we suppressed
by excluding data lying outside 5 MAD's of the median. No distributions were sufficient to
cleanly fit the full set of data, so we focus here on the center of the distribution and include
a qualitative discussion of the heavy tails below. With the exception of ∆G+
mech which was fit
with a generalized normal distribution, each distribution exhibited significant skew and could
be fit reasonably well with a log-normal or a gamma distribution. At higher concentrations of
myosin and cross-linkers, the histograms were less cleanly fit by any standard distributions (SI
Figure 2).
Log-normal distributions are fairly ubiquitous across different fields and systems [46].
For instance, it has been shown that the distributions of concentrations of species in a chemical
reaction network are log-normal [47]. Gamma distributions are similarly common and often
difficult to discriminate from log-normal distributions [48]. A speculative explanation for the
gamma distribution of ∆Gstress is as follows: ∆Gstress can be viewed as resulting from a number
of myosin filament steps that, according to the central limit theorem, is approximately normally
distributed given a sufficiently long time between simulation snapshots tsnap (these stepping
events are not truly i.i.d., but to a first approximation we may assume they are). The main effect
of each of these steps is to increase the harmonic stretching potential on the myosin filaments as
well as on the actin filament cylinders by a roughly fixed stepping distance. Thus the increase in
mechanical energy, ∆Gstress, is approximately a quadratic function of the normally distributed
number of myosin filament steps per tsnap. As shown in SI Figure 2, the resulting distribution
of ∆Gstress is well-fit by a gamma distribution and bears qualitative similarity to the histogram
of ∆Gstress in Figure 5. In Figure 5, ∆Gstress and ∆G+
mech, dissipated have similar distributions,
indicating that, for each MEDYAN cycle, almost all the stress accumulated following Step 1
is then immediately relaxed. The remainder goes into ∆G+
mech, whose distribution is centered
10
on zero with little skew. Furthermore, the distribution of ∆G+
mech, dissipated has particularly
heavy tails, indicating infrequent yet large relaxation events. This tendency has some precedent
in avalanche-prone systems, whose hallmarks are self-organized criticality, intermittency, and
scale-invariance in their distribution of avalanche event sizes [49, 50, 51, 52]. While we do
not observe true scale-invariance in this set of simulations (i.e. power-laws cannot fit these
distributions cleanly), it will be interesting to continue to explore actomyosin networks dynamics
in the framework of self-organized criticality [53].
3.2 More Compact Networks are More Efficient
dissipated, ∆G+
We simultaneously varied concentrations of myosin filaments CM , and cross-linkers, CCL, which
is known to produce a range of network architectures [34, 54]. Using cross-linker concentrations
of 0.1, 1.0, and 5.0 µM , and myosin concentrations of 0.1, 0.2, and 0.4 µM , we studied the
effects on ∆G+
mech, dissipated. The concentrations of actin
subunits and filaments are the same as described above. For each of the 9 conditions, we ran
10 simulations of 2000 s. In Figure 6, we display the median values of ∆G+
chem, dissipated and
∆G+
mech, dissipated along each trajectory and over each repeated trajectory for these conditions.
The median is used here because of its insensitivity to outliers such as are found in the heavy-
tailed distributions of these quantities [42].
chem, dissipated, and ∆G+
Figure 6: Bar plot representing the contributions of ∆G+
mech, dissipated to
the total, ∆G+
dissipated. The letters in the abscissa labels designate "low," "medium," and "high."
The first letter represents the concentration of myosin, CM : "L" = 0.1 µM , "M" = 0.2 µM ,
"H" = 0.4 µM , and the second letter represents the concentration of cross-linkers, CCL: "L" =
0.1 µM , "M" = 1.0 µM , "H" = 5.0 µM . The median of each quantity is taken over 10 runs of
2000 s, and error bars represent 1 MAD.
chem, dissipated and ∆G+
11
chem, dissipated increases. As CCL is increased with CM fixed, then ∆G+
We find that as CM is increased, the median rates of ∆G+
mech, dissipated all tend to increase. Further, the value of of ∆G+
chem, dissipated,
and ∆G+
mech, dissipated relative to
∆G+
chem, dissipated tends
to decrease. Increased concentrations of myosin filaments obviously have a strong effect on the
dissipation simply because they are the chief active agents in the system once treadmilling has
reached its steady-state. We can define a measure of efficiency for the present purpose as:
dissipated, ∆G+
η =
∆Gstress
∆G+
chem
=
∆G+
mech, dissipated − ∆G+
∆G+
dissipated − ∆G+
mech
mech
≈ ∆G+
mech, dissipated
∆G+
dissipated
(6)
where the approximation follows since, as illustrated in Figure 2, the average of ∆G+
mech is
close to zero. Thus it is evident that, perhaps surprisingly, as CM increases, η increases: the
more motors are in our system, the more efficiently can chemical energy be converted into
mechanical stresses. Further, as CCL increases, η tends to increase because ∆Gchem, dissipated
decreases relative to ∆Gmech, dissipated. At higher levels of cross-linking, which introduce me-
chanical constraints on the filaments, the walking of myosin motor filaments will produce more
stress compared to at low levels of cross-linking, when filament sliding can result from myosin
filament walking, producing less stress.
We repeated the above experiments with the inclusion of mechanochemical feedback on
the reaction rates controlling cross-linker unbinding and myosin filament unbinding and walking
(SI Figures 2). We observed, somewhat surprisingly, no qualitative differences compared to the
results described for the case of no feedback, however there were quantitative differences in the
stress production and radius of gyration for certain conditions. These quantitative differences
result from the fact that, for the concentrations CM and CCL used above, we did not observe
significant collapse of the actomyosin network when mechanochemical feedback was included due
to the stalling and catching of the myosin filaments. As a result the networks were less densely
cross-linked, and the efficiency was lower. The total dissipation rates are largely unaffected
by the inclusion of mechanochemical feedback; it was primarily the degree to which chemical
energy had been converted to stress that was different for certain conditions.
It is worthwhile to mention how these results compare with in vitro studies of dissipa-
tion in actomyosin systems. While the computational approach described in this paper allows
uniquely highly-resolved and direct measurement of free energy changes, other experimental
methods have produced qualitatively similar results to those obtained here. Rheological ex-
periments have determined that a single cell's response to compression is similar in nature to
a muscle's response to increasing load, suggesting that the actomyosin network underlies the
cell's mechanical responsiveness [31]. Further, this responsiveness is modulated by blebbistatin,
a myosin ATPase inhibitor, highlighting myosin's role in negotiating how the network rearranges
in response to the sustained stress. Using the metric of mechanical dissipation to measure the
degree of structural rearrangements and release of stress, we confirm that these processes indeed
sensitively depend on myosin activity, which we control here through its concentration. Addi-
tional rheological studies have probed the mechanical dissipation of actin cortices more directly,
using the loss modulus as a readout. These have also indicated that inhibiting myosin activity
reduces the mechanical dissipation of the system, causing it to be behave more elastically [30].
Lastly, we mention a recent study that quantified mechanical dissipation of actin filaments us-
ing a novel experimental method [55]. By measuring the flow through a low-dimensional phase
space defined by the amplitudes of the filament bending modes [56], they determine the entropy
production of fluctuating actin filaments in different phases of contractility. They relate the en-
tropy production to the degree of transverse bending of the filaments, as opposed to sarcomeric
filament sliding, caused by myosin filament walking. Similarly, we here relate the hindrance of
12
filament sliding due to cross-linker density to increased mechanical dissipation rates. Quanti-
tative comparisons of these two approaches to mechanical dissipation measurement will be an
interesting future direction. We note finally that a unique capability of quantifying dissipation
using MEDYAN is the ability to simultaneously measure the energetics of chemical reactions in
addition to the changes in the mechanical energy, which is not currently available using in vitro
methods.
4 Discussion
We have introduced a methodology for tracking the energetics of chemical and mechanical events
during a MEDYAN simulation, allowing us to probe the properties of actomyosin networks
as dissipative active matter systems. The distinction between dissipation's mechanical and
chemical origins is natural in the context of MEDYAN's iterative simulation procedure which
carries out chemical stochastic simulation, mechanical deformation, and mechanical relaxation
at separate times. As explained in [34], this procedure exploits a separation of timescales
between the characteristic mechanical relaxation times of actomyosin networks [57] compared
to typical waiting time between reactions that introduce mechanical stresses, such as myosin
filament walking [10] or filament growth [58]. Ultimately the source of all dissipation is the
chemical potential of ATP molecules driving treadmilling and myosin filament activity. This is
reflected by the near equality of ∆G+
dissipated in Figure 2 and SI Figures 2. On a
fast timescale (that of the characteristic mechanical relaxation time), however, the free energy
of chemical reactions cause small force deformations of the network which are then quickly
and almost fully relaxed. Thus, for a myosin filament stepping event, only a portion of the
chemical free energy ∆G+
chem, dissipated is immediately dissipated as heat, with the rest going
into temporarily increased mechanical energy of the actomyosin network, ∆Gstress. The fast
relaxation of this new mechanical stress constitutes what we refer to as mechanical dissipation,
∆G+
mech, dissipated, and the small residual stress after this relaxation has balanced all net forces
acting on the system results in a change of the mechanical energy of the system on a slow
timescale, ∆G+
chem and ∆G+
mech.
One interpretive framework which is useful to understand the flow of free energy in
actomyosin networks is illustrated in Figure 7. We can think of the different forms of free energy
storage, including chemical potentials, mechanical stress, concentration gradients (which could
also be considered as arising from chemical potential differences across compartments), and
dissipated energy, as nodes on a directed graph, where edges represent transduction of energy
from one form of storage to another. The weights of these edges represent the amount of
free energy flowing through them.
In this picture, we can describe the process of network
percolation, which occurs at increasing concentrations of cross-linkers and myosin filaments,
as widening the edge flowing from chemical potential into mechanical stress, while thinning
the edge from chemical potential directly to dissipation. The edge weights corresponding to
the establishment of concentration gradients and the resulting diffusive dissipation will not be
affected dramatically by the onset of percolation except to the extent that percolated networks
might lead to the formation of more bundles, and therefore more significant concentration
gradients. At steady state, we have a stationary current on the graph, fueled by the chemical
potential of the assumed limitless supply of ATP. Of course, at this level of description, we
have coarse-grained away the details of the specific chemical reaction networks and mechanical
potentials which constitute the system, however by doing so we gain a clearer understanding of
how percolation of the network alters the flows of free energy.
13
Figure 7: A simple schematic illustrating the flow of free energy in actomyosin network systems.
Blue compartments represent forms of free energy storage, arrows represent the transfer of free
energy from one form to another, and the green compartment indicates dissipation as ultimate
destination of all free energy flows in this nonequilibrium system. Arrows are labeled with the
mechanisms by which these free energy transduction from one form to another are achieved. In
this depiction, the sizes of compartments and the widths of the arrows indicating the magnitudes
of the represented quantities are not to scale.
The new capabilities of MEDYAN allowing detailed energetics computations should
provide a way to address outstanding issues in the the fields of active matter systems and of
cell mechanics. In this study we observe dissipation rates to stay at a low fluctuating steady-
state following a high initial transient phase, apparently in contrast with the predictions of the
dissipation-driven adaptation hypothesis. However we also do not observe slow reorganization
of the actomyosin network into different higher order structures such as bundles (for the concen-
trations of components used here we observe disordered networks only), and as a result we can
not rule out that such reorganizations correspond to marked changes in the dissipation rates. A
dedicated test of the hypothesis of dissipative-driven adaptation should be straightforward with
this methodology. For instance, a simulated gliding assay, which has been shown in vitro to lead
to diverse dynamical patterns [20], may indicate that these emergent patterns correspond to an
optimization of the chemical dissipation from myosin filament walking, as argued in [25]. In the
context of cell migration, studies investigating the mechanically dissipative activity of myosin
filaments are also feasible if one incorporates in simulation an external substrate against which
the actomyosin network pulls. In this type of study, it should be straightforward to determine
how the amount of stress which is sustained against the substrate is altered by myosin activity
given the energetics calculations in the methodology described in this paper. It is also feasible
to do simulated measurements of the dynamic shear modulus by compressing the actomyosin
network at different frequencies. It should then be possible to directly observe the degree to
which the dissipation of elastically stored energy is attributable to myosin walking. In fact, this
last research question has already been investigated to some extent using an alternate computa-
tional model to the one presented here [59]. We hope some of these potential future experiments
14
Chemical PotentialMechanical StressConcentration GradientsDissipationMyosin WalkingTreadmilling Heat of ReactionsMechanical RelaxationDiffusionwill shed light on outstanding questions in the studies of actomyosin networks.
5 Supplementary Information for "Quantifying Dissipation in
Actomyosin Networks"
5.1 ∆G of Chemical Reactions
For a reaction of the general form
ν1X1 + ν2X2 + . . . (cid:10) υ1Y1 + υ2Y2 + . . .
(7)
we define Xi as the reactant species, Yi as the product species, and νi and υj as their stoichio-
metric coefficients. The "stoichiometric difference" is defined as
σ =
νi
(8)
(cid:88)
j∈P
υj −(cid:88)
i∈R
where P is the set of products and R is the set of reactants. We further define the conversion
factor between the copy number of species i, Ni, and its concentration Ci,
Θ = NAvV
(9)
where NAv is Avogadro's number, V is the volume of the compartment where the reaction
occurs, and we have Ni = CiΘ. Next we define the following quantity, reminiscent of the
reaction quotient:
(cid:89)
(cid:101)Q =
(cid:89)
(Ni − νi)Ni−νi
(Nj + υj)Nj +υj
i∈R
N Ni
i
j∈P
N Nj
j
where R is the set of reactant species and P is the set of product species. Lastly, defining ∆G0
as the standard state change in Gibbs free energy, we arrive at the following expression for the
change in Gibbs free energy as a function of the instantaneous vector of species copy number
N:
∆G(N) = ∆G0 − σkBT log Θ − σkBT + kBT log (cid:101)Q.
(11)
Although this equation might appear unusual, upon making some approximations leveraging
the relative size of the stoichiometric coefficients and the species copy numbers it can be shown
to reduce to the familiar textbook expression
(10)
(12)
(13)
where
Q =
∆G(C) = ∆G0 + kBT log Q
(cid:89)
i∈R
(cid:89)
j∈P
−νi
i
C
Cυj
j
is the reaction quotient of the species concentrations Ci, and C is the vector of these concen-
trations. In the context of a compartment-based reaction diffusion scheme, the species copy
numbers used in Equation 11 are specific to the compartment in which the reaction was drawn.
We prefer to use Equation 11 in simulation because small deviations from this nearly exact
expression can actually lead to significant systematic bias of the change in Gibbs free energy
resulting from certain reactions. This is especially true for reactions that are very frequent,
such as the diffusion reaction between adjacent compartments. ∆G for diffusion reactions, in
which a molecule jumps from a compartment where its concentration is Ni,A to a compartment
where its concentration is Ni,B, is calculated using an expression similar to Equation 11:
∆G = kBT log
(Ni,A − 1)(Ni,A−1)
(Ni,B + 1)(Ni,B+1)
N Ni,A
i,A
15
N Ni,B
i,B
.
(14)
Similarly this expression reduces to the familiar formula
∆G = kBT log
Ni,B
Ni,A
(15)
upon leveraging the sizes of Ni,A and Ni,B relative to 1. The derivation of the above results can
be found in an accompanying paper [38].
For some of the reactions in the CT model, Equation 12 can be directly applied and
straightforwardly cast into the form of Equation 11 to give more exact results. However other
reactions including (de)polymerization, myosin filament walking, myosin filament (un)binding,
and cross-linker (un)binding, require some additional treatment.
For (de)polymerization reactions, care should be taken in defining the concentrations
of the reactants and products since those molecules include heteropolymers; we would like
to avoid requiring that a specific sequence of distinct subunits constitutes a unique chemical
species. This is possible to do because the polymerization of actin subunits is independent of
the chemical identity of the subunit at the tips of the polymer, and is therefore independent of
the chemical identity of any of the polymerized subunits [39]. Then, polymerization potentially
only depends on the polymer length, such that a general reaction for reversible polymerization
can be written as
(m)n + m ←→ (m)n+1
(16)
where (m)n is a polymer with degree of polymerization n and m is a subunit. For this reaction
Equation 12 reads
∆G = ∆G0 + kT log
.
(17)
C(m)n+1
C(m)nCm
Following [60], we make the simplifying assumption that a polymer's chemical reactivity does
not depend on its degree of polymerization, provided that the polymer is sufficiently long and
cooperative effects do not apply. With this, Equation 17 reduces to
∆G = ∆G0 + kT log
1
Cm
.
(18)
This assumption agrees more or less with intuition: the polymerizing subunit does not really
"see" the degree of polymerization of the polymer, which is reflected by constant rates of
polymerization for polymers of varying lengths.
Cross-linking proteins are incorporated in MEDYAN as diffusing species that bind and
mechanosensitively unbind to pairs of actin filaments, mechanically coupling them. We treat
the change in free energy of the (un)binding reactions analogously to the (de)polymerization
reactions. Ignoring the chemical identities of the heteropolymer subunits, a general cross-linking
protein (un)binding reaction can be written as
(m)i + (m)j + L ←→ (m)iL(m)j
(19)
where L is a cross-linking protein and (m)i is a polymer with degree of polymerization i.
According to Equation 12, the change in Gibbs free energy upon this reaction occurring to the
right is
∆G = ∆G0 + kT log
.
(20)
C(m)iL(m)j
C(m)iC(m)j CL
Similarly to our assumption that the chemical reactivity of polymers is independent of degree
of polymerization for sufficiently long polymers, we assume here that the binding affinity of
cross-linking proteins is independent of the degree of polymerization and of the number of
cross-linking proteins already bound to the pair of filaments, allowing us to simplify Equation
20 to
∆G = ∆G0 + kT log
16
1
CL
.
(21)
The kinetics of myosin filament walking in MEDYAN are based on the Parallel Cluster
Model (PCM) [61]. Taking the rate constants of individual myosin head (un)binding reactions
as inputs, and accounting for the statistical distribution of the number of bound heads as well
as dependence on the force exerted on the filament, the results of the PCM allow us to write
kinetic parameters describing the entire myosin filament, including filament (un)binding and
walking rates. In the MEDYAN implementation, each step of a myosin filament can represent
several steps of the constituent myosin heads, where each head step represents the completion of
a single myosin head cross-bridge cycle [12, 62]. The head steps have a fixed length, dstep, set by
experimental measurements of the myosin isoform of interest. The length of the filament steps,
dtotal, is determined from the following MEDYAN parameters: the equilibrium length Lcyl of the
cylinders that comprise the coarse-grained representation of actin filaments, and the number
of binding sites per cylinder Nbs, giving dtotal = Lcyl/Nbs. The binding sites represent the
discrete locations on the cylinders which can be occupied by cross-linkers or myosin filaments.
To account for the discrepancy between dstep and dtotal, we multiply the filament walking rate
by the ratio
(22)
When a filament step occurs in MEDYAN, it thus represents the completion of s−1 myosin head
cross-bridge cycles, each of which has the effect of converting one solvated ATP molecule into
solvated Pi and ADP. The Gibbs free energy the filament walking reaction is then
s =
dstep
dtotal
.
(cid:18)
(cid:19)
CADP CP i
CAT P
∆G = s−1
∆G0 + kBT log
.
(23)
where ∆G0 refers to the standard change in Gibbs free energy for hydrolysis of ATP. This
expression for ∆G could be further multiplied by a parameter ζ representing the coupling of the
ATP hydrolysis cycle to the forward step of the myosin head; here we follow the assumption of
tight coupling, i.e. ζ = 1 [63, 62].
The mechanochemical updating of force-sensitive reaction rates k± alters the equilib-
rium constant via Keq = k−/k+, and therefore the change in free energy ∆G corresponding to
that reaction via ∆G0 = kBT log Keq. For a reaction whose rates have been updated due to
some applied force F , it can be shown that the new value of ∆G0 is approximately given by
∆G0(F ) = ∆G0F =0 + g(F )
(24)
where G0F =0 is the original (i.e. zero-force) value of ∆G0 and g(F ) is equal to the increase
mechanical energy due to the applied force [7, 62]. In this modeling, the extra energy g(F ) is
counted as a part of ∆Gmech, dissipated, not ∆Gchem, dissipated. So when, for example, a cross-
linking protein unbinds under tension F , the zero-force value ∆G0F =0 is used when computing
the change in free energy for that reaction, and when mechanical equilibration next occurs, the
released stretching energy g(F ) is included in the calculation of ∆Gmech, dissipated.
5.2 Parameterization
Parameterization of the CT model for the purpose of tracking free energy changes during sim-
ulation trajectories consists of choosing values of the rate constants (kinetic parameters) and
of ∆G0 (thermodynamic parameters) for all reactions in the model. Wherever possible, values
from the literature are used. Experimental measurements have determined rate constants for
every reaction, however for some reactions the value of ∆G0 hasn't been reliably measured, to
the best of our knowledge. Below, we describe a technique to solve for these unknown values.
For reversible reactions, where the forward and reverse rate constants k+ and k− are
known, such as (de)polymerization and (un)binding of cross-linkers, ∆G0 can be found from
∆G0 = kBT log Keq = kBT log
17
k−
k+
(25)
Literature values for the equilibrium constants or of ∆G0 are used for irreversible reactions,
for which k− is often too small to determine from direct measurement. Irreversible reactions
in this system include all reactions except for (de)polymerization reactions. For reactions for
which literature values of ∆G0 are unavailable, it is possible to solve for ∆G0 values based on a
self-consistency condition [64, 65]: the sum of the ∆G0 values around a closed loop of reactions
in which the number of molecules has not experienced a net change must be zero since the free
energy is a state function (equivalently, by Equation 25, the product of equilibrium constants
around any such loop must be equal to one). Writing several such closed loops of reactions
leads to a system of equations that can be solved for the unknown variables. Not all possible
loops result in independent equations, but we were able to determine the values of two unknown
parameters using the loops illustrated in SI Figure 8.
Figure 8: Diagrams representing sequences of reactions, with the involved species drawn be-
tween each reaction, resulting in independent relations between the equilibrium constants. The
meaning of these equilibrium constants is provided in the main text. Loops are assumed to
proceed in the clockwise direction, and arrows that point opposite to this direction indicate
that that reaction is occurring "backwards", i.e.
from products to reactants. Polymers are
shown as connected chains of subunits, while the "+" sign indicates different solvated species.
The loops in SI Figure 8 imply the following independent system of equations:
KT KhydKrelKnex = KP iKATP
KT KhydKphosKnex = KDKATP
(26)
(27)
where Khyd represents the hydrolysis of ATP by F T , Kphos represents the release of Pi by F P i,
Knex represents nucleotide exchange converting GD to GT , KATP represents the hydrolysis of
ATP in solution producing ADP and Pi, and Krel represents the release of phosphate by GP i.
18
...+.........+...++...++++++++KTKhydKPiKrelKnexKATP...+.........++...+++++++KTKhydKDKnexKATP...+Kphos+ATP ActinADP-Pi ActinADP ActinATP PiADPA)B)Values from the literature [58, 13, 11] can be used to determine 6 of the 8 variables
in Equations 26 and 27, which thus represent two equations in two unknowns: Krel and Knex.
The resulting parameters are listed in Table 1.
Note that it is possible to draw loops such as those in SI Figure 8 that would imply
that the equilibrium constants for polymerization and depolymerization of, for example GT ,
should be the same at the plus and minus ends of the filaments. This condition is not borne out
by the experimental values of these equilibrium constants, and this discrepancy is a recognized
outstanding problem [58]. Here, we use the literature values for these equilibrium constants and
employ the reaction loop method only to determine the parameters Krel and Knex.
For reversible binding of myosin filaments, results from the PCM are used to describe
binding and unbinding rates, and therefore Keq [61, 12]. The filament binding rate is given as
and the unbinding rate is
kfil,unbind = Ntkhead, bind
kfil, bind = Ntkhead, bind,
khead, unbind
(cid:21)−1
(cid:20)(cid:18) khead, bind + khead, unbind
(cid:19)Nt − 1
(cid:19)Nt − 1
(cid:20)(cid:18) khead, bind + khead, unbind
(cid:21)
,
khead, unbind
where khead, bind and khead, unbind describe the binding kinetics of a single myosin head, and Nt
is the number of heads in the filament [61]. The resulting expression for ∆G0 is
∆G0 = kT
kfil, unbind
kfil, bind
= −kT log
.
(30)
We assume that we have chemostatted concentrations of ATP, ADP, and Pi, which we
account for implicitly via the effect of these concentrations on the kinetic and thermodynamic
parameters of certain reactions. Thus these species are not explicitly tracked. The concen-
trations of these species affect the change in Gibbs free energy associated with the following
reactions:
• myosin filament walking
• nucleotide exchange (GD → GT )
• phosphate release by F and G-actin (F P i → F D, GP i → GD)
The effect of the concentrations of ATP, ADP, and Pi for the these reactions is to simply
change the reaction quotient Q which changes ∆G via ∆G = ∆G0 + kBT log Q. For instance
the nucleotide exchange reaction can be written explicitly as
AT P + GD → ADP + GT
(31)
and the change in Gibbs free energy is
∆G = ∆G0 + kBT log
(cid:18) CADP CGT
(cid:19)
CAT P CGD
.
To treat the concentrations of ATP and ADP implicitly, we rewrite the reaction as
GD → GT
for which the change in Gibbs free energy is
∆G = ∆G0(cid:48)
+ kBT log
where
∆G0(cid:48)
= ∆G0 + kBT log
19
(cid:19)
(cid:18) CGT
(cid:19)
(cid:18) CADP
CGD
,
CAT P
.
(28)
(29)
(32)
(33)
(34)
(35)
A similar approach is taken for the other reactions mentioned above.
The concentration of just ATP (since neither ADP or Pi appear implicitly as reactants
in any of the reactions of the CT model) affects the kinetics of the following reactions:
• myosin filament walking
• nucleotide exchange
To understand the effect of CAT P on the myosin filament walking rate, we employ the five state
cross-bridge model of a single myosin head described in [12]. In that model unbinding of a head
from the filament substrate occurs via two pathways: a slip path, with rate k35, and a catch
path, with an effective rate k345. The catch path is a two-step reaction: the release of ADP
with rate k34, followed by the unbinding of the filament head and binding of ATP with rate
k45 = kT CAT P . The effective rate constant of the approximate one-step representation of this
reaction is
k345 =
.
(36)
k34kT CAT P
k34 + kT CAT P
Unless CAT P is very low, this reaction rate is limited by k34. Because the head can unbind by
the catch or slip pathway, the rate for unbinding is
khead, unbind = k35 +
k34kT CAT P
k34 + kT CAT P
(37)
The slip path should also be dependent on ATP concentration, since in state 5 the head is
ATP-bound, however we follow the authors of [12] in neglecting this dependence since the slip
path only becomes active under large load.
The rate of the nucleotide exchange reaction also depends on CAT P and also occurs in
two steps. The full reaction, with ATP and ADP explicitly included, is
GD + AT P → G∗ + AT P + ADP → GT + ADP
(38)
where G∗ represents actin with no bound nucleotide. Following [39], this reaction is approxi-
mated as a one-step irreversible reaction with ATP and ADP included explicitly, Equation 33.
We can write the rate knex of this approximate reaction as:
knex =
kD→∗k∗→T CAT P
kD→∗ + k∗→T CAT P
≈ kD→∗
(39)
where kD→∗ is the dissociation rate of ADP, k∗→T is the second order rate constant of ATP
association, and the approximation holds except at low concentrations of ATP when ADP
dissociation is no longer the rate-limiting step. The values of k∗→T and kD→∗ are presented and
discussed in [66, 67, 68].
20
Reaction
GT poly at plus end
GT depoly at plus end
GT poly at minus end
GT depoly at minus end
GP i poly at plus end
GP i depoly at plus end
GP i poly at minus end
GP i depoly at minus end
GD poly at plus end
GD depoly at plus end
GD poly at minus end
GD depoly at minus end
Pi release by F-actin
ATP hydrolysis by F-actin
Pi release by G-actin
Nucleotide exchange
Cross-linker binding
Cross-linker unbinding
Myosin head binding
Myosin head unbinding
Myosin filament walking
Rate Constant ∆G0 (kBT ) Reference
11.6 (µM s)−1
1.4 s−1
1.3 (µM s)−1
0.8 s−1
3.4 (µM s)−1
0.2 s−1
0.11 (µM s)−1
0.02 s−1
2.9 (µM s)−1
5.4 s−1
0.09 (µM s)−1
0.25 s−1
0.002 s−1
0.3 s−1
5 s−1
0.01 s−1 e
0.7 (µM s)−1
0.3 s−1
0.2 s−1 d
1.708 s−1 d
- f
-2.12 a
2.12 a
-0.51 a
0.51 a
-2.81 a
2.81 a
-1.75 a
1.75 a
0.59 a
-0.59 a
1.03 a
-1.03 a
-7.31 b,e
-10.0
-10.77 c,e
-6.76 c,e
-0.85 a
0.85 a
- d
- d
-14.5 e
[39]
[39]
[39]
[39]
[58]
[58]
[58]
[58]
[39]
[39]
[39]
[39]
[39, 58]
[39, 13]
-
[39]
[34]
[34]
[34]
[34]
[69]
Table 1: Kinetics and thermodynamic parameters describing reactions in the CT model as well
as cross-linker and myosin filament (un)binding and myosin filament walking.
a - Values of ∆G0 determined via Equation 25.
b - ∆G0 determined from Keq given in [58].
c - ∆G0 determined using constraints as described above.
d - Parameters describing the myosin filament obtained via Equations 28, 29, 30.
e - Depends implicitly on CAT P , CADP , and CP i; given value applies to standard state.
f - Calculated in simulation using results of PCM, see [34].
In all the studies in this paper, implicit nucleotide concentrations are taken to be CAT P
= 8 mM , CADP = 7 µM , and CP i = 1 mM , corresponding roughly to the amounts found
in human muscle after exercise [69]. Other parameters of the system, including mechanical
constants, diffusion rates, screening lengths, and boundary cutoffs have been set to the same
values given and discussed in [34]. Cylinder equilibrium lengths Lcyl are chosen as 27 nm with 4
binding sites per cylinder for myosin filaments and 1 binding site per cylinder for cross-linkers,
giving approximately physiological values for stepping distances of myosin motor filaments and
spacing along actin filaments of bound α-actinin. We note that the form of the mechanochemical
models has been changed from those used in [34]; the modeling used here is current as of
MEDYAN v3.2, and we refer readers to the documentation at http://www.medyan.org/ for
further details.
5.3 Mean-Field Model of Treadmilling Dissipation
To validate the methods for quantifying dissipation using MEDYAN against a simpler represen-
tation of actin filament treadmilling, we developed a mean-field description of the dissipation
resulting from chemical reactions in the CT model. Mean-field models of the trajectory of the
vector of concentrations of species, C(t), have been formulated previously as an 11-dimensional
system of ordinary differential equations (ODEs) ([39]), and in the CT model as a 5-dimensional
21
system of ODEs ([40]). These models describe the polymerization of a concentration Nfil of actin
filaments in a pool of actin subunits of total concentration M . The subunit species tracked by
these models are distinguished by their polymerization state and by the hydrolysis states of the
nucleotide to which they are bound. The meaning of "mean-field" in this context is the as-
sumption that reacting species are well-mixed over the entire system volume, therefore obeying
mass-action kinetics and deterministic dynamics which can be represented by ODEs. For an
instantaneous value of C, we define a function DΛ(C) representing the instantaneous rate of
dissipation due to a set of reactions Λ. Thus a solution of a mean-field C(t) = N(t)/Θ, allows
us to construct the trajectory of the dissipation rate, DΛ(C(t)). This function cannot capture
the dissipation due to the activity of myosin filaments or cross-linkers, or from relaxation of
mechanical stress, because these aspects are not included in the these mean-field models de-
scribing filament treadmilling. The benefit of such a mean-field model is that one can perform
systematic variation of parameters with limited computational demands, and we use it here to
study the effect of the parameters Nfil and M on the dissipation due to filament treadmilling.
The function DΛ(C(t)) can be written as a sum over the reactions λ ∈ Λ of the
instantaneous rate of change of the solution's Gibbs free energy due to that reaction:
DΛ(C(t)) =
∆Gλ(C(t))rλ(C(t)).
(40)
(cid:88)
λ∈Λ
(cid:89)
i∈R
The expressions for ∆Gλ(C(t)) for different reactions are described above. The instantaneous
rate of reaction λ, rλ(C(t)), is written as usual for mass-action kinetics as
rλ(C(t)) = Θkλ
Cνi
i
(41)
where R is the set of reactant species for reaction λ, and where we have included the conversion
factor Θ to convert rλ(C(t)) to units of s−1.
To facilitate the study of dissipation due to filament treadmilling, we define a set of
reactions Λ in the CT model which constitutes the dominant cycle a subunit undergoes in the
treadmilling process: a) polymerization of GT to the plus end, b) hydrolysis of ATP by F T ,
c) release of Pi by F P i, d) depolymerization of GD from the minus end, and e) nucleotide
exchange converting GD to GT . We refer to this set of 5 reactions as the main treadmilling
pathway (MTP). Alternative sequences of reactions whose net effect is similarly the conversion
of one molecule of ATP to ADP and Pi are considered as of secondary importance and not
included in this analysis.
This mean-field description of entropy production can be compared to the description
of entropy production rates for chemically reactive systems that emerged from the Brussels
school of thermodynamics. [70, 71]. The results of that school include the minimum entropy-
production principle applicable in the linear regime [22], and the general evolution criterion
applicable even in the the non-linear regime [72]. Their formalism typically considers the total
entropy production rate as a volume integral over the local entropy production rate density,
which is itself written as a sum over fluxes multiplied by their corresponding thermodynamics
forces defined at each point of the system. This sum is decomposed into terms representing
diffusion and terms representing chemical reactions, and the terms representing the chemical
reactions are written such that the fluxes reflect the net reaction rate at that point in the system.
In contrast, our mean-field description neglects concentration gradients and resulting diffusion
fluxes, as we assume a homogeneous distribution of the chemical species. Equation 40 then
represents only the chemical contribution to the entropy production, as a sum over the rates
and affinities of the reactions in the system, implicitly integrating over the homogeneous system
volume. We treat the forward and reverse direction for some chemical reaction as separate terms
in Equation 40, so these rates cannot be considered fluxes which would include the reverse rate as
well. This allows for more general sets of reactions which might include effectively irreversible
22
processes for which the reverse rate is negligible. However the set of reactions Λ could be
chosen to include a reverse reaction for each forward reaction, with the result that these pairs
of terms represent fluxes along the reversible reaction pathway. The parsimonious reaction set
MTP is chosen not to fully describe the rate of entropy production in the system, but to allow
easy visualization of the main contributions to the entropy production. We do not pursue the
connection of our treatment to the formalism of the Brussels school further here, but we lastly
note that our results are compatible with their minimum entropy-production principle, as shown
in SI Figure 9.
We first verified that the mean-field model of MTP dissipation agreed with results
from MEDYAN simulations, to illustrate consistency between these approaches. In SI Figure 9
we display the close match between the trajectory of MTP dissipation over a 2000 s run from
these two approaches. Note that, to allow direct comparison, only the changes of Gibbs free
energy resulting from reactions in the MTP set are visualized for both approaches here, i.e.
the contribution from diffusion and other non-MTP reactions in the MEDYAN simulation are
not visualized. We also turned off in MEDYAN the force-sensitive decrease in polymerization
rate when the filament tips push against the simulation hard-wall boundaries, since this effect
is not captured in the mean-field modeling [73]. We used a simulation volume of 1 µm3 and
initial conditions of equal amounts (10 µM each) of GT and GD actin in a 0.08 µM pool of
seed filaments containing F T . The dissipation rate decreases nearly monotonically, attaining a
minimal steady-state value after tens of seconds. In SI Figure 9 we also display the individual
contributions to the sum in Equation 40. Initial polymerization of GT to the plus end constitutes
the majority of the initial dissipation. As this polymerization process slows after about 1 second,
the hydrolysis of ATP by the now relatively abundant F T becomes the dominant contribution.
As hydrolysis then slows after about 10 seconds, the total dissipation rate reaches a steady-
state value of roughly 80 kBT /s. In SI Figure 9 we also plot the mean-field prediction of the
trajectory of the reacting species' concentrations.
23
Figure 9: Results of the mean-field modeling of MTP dissipation. Top: Comparison of
DMTP(C(t)) calculated using the mean-field model (with time increments of 0.01 s), with
∆Gchem rates measured during MEDYAN simulation (with time increments of 0.1 s). Mid-
dle: Plot of the contributions of each reaction in the MTP to the total dissipation. The items
in the legend represent reactions in the MTP, which are described in the main text. Bottom:
Plot of the trajectories of the reacting species concentrations. The notation for each species is
described in the main text.
We next simultaneously varied the total concentration of actin M and the concentra-
tion of actin filaments Nfil, and determined the total dissipation integrated along each trajectory
as well as the steady-state dissipation rate. As M was varied, we held the initial concentration
of each species proportionally the same: 49 % GT , 49 % GD, and 2 % F T . As shown in SI Figure
10, the integrated dissipation over 2000 s was observed to increase monotonically with both M
and Nfil. Quantitatively, the integrated dissipation depends on the choice of initial proportions,
24
however we found that the shape of the dependence on Nfil and M is largely independent of
initial proportions (data not shown). Total dissipation increases with M simply because more
actin is available to hydrolyze ATP during the approach to steady-state. For large amounts of
actin, increasing Nfil, the number concentration of filaments, allows increased polymerization
of GT , which constitutes a large contribution to total dissipation during the early stages of the
trajectory. Increasing Nfil also shifts the steady-state concentration of GT downward (SI Figure
11), implying that more GT has been polymerized during the approach to steady-state. A loose
analogy can be made of a one lane road compared to a multi-lane highway during heavy traffic
to describe this situation. As Nfil is increased with M fixed, the steady-state dissipation rate
increases concavely, as shown in SI Figure 10. The steady-state concentration of F T also in-
creases concavely (SI Figure 11), representing higher rates of ATP hydrolysis. The contributions
of each reaction to the total dissipation rate at steady-state as Nfil is varied is illustrated in SI
Figure 12. The lack of dependence of the steady-state dissipation rate on M can be explained
by the fact that increasing M increases the steady-state concentration of only F D, not of any
other species [40]. In other words, all extra actin accumulates in the form F D as M is increased.
This species is essentially inert, since the depolymerization rate of F D is controlled by the con-
centration of filaments Nfil. Thus the steady-state dissipation has no dependence on the total
amount of actin. Furthermore, it has no dependence on the initial concentrations, since it is
known that the steady-state vector of concentrations does not depend on initial conditions [40].
Figure 10: Left: The total dissipation integrated over 2000 s trajectories as M and Nfil are
varied. Right: The steady-state dissipation rate over the same range of M and Nfil.
25
Figure 11: The concentrations at steady-state of the various actin subunit species as the con-
centration of filaments Nfil is varied. These curves have no dependence on initial conditions,
except F D which will increase linearly with the total concentration of actin subunits M ; any
additional actin subunits in the system will accumulate in the form F D at steady-state.
26
Figure 12: Contributions of each reaction in the main treadmilling pathway to the total dissi-
pation rate at steady-state, as the concentration of filaments Nfil is varied. These curves have
no dependence on the initial concentrations of the different subunit species.
Data accessibility
MEDYAN is available under copyright for download at http://www.medyan.org/. The version
code and data analyzed here without mechanochemical feedback are available through the digital
repository at the University of Maryland: https://doi.org/10.13016/t8oa-9qra.
Authors' contributions
G.P. conceived the project. G.P. and C.J. consulted on all aspects of the project. C.F. developed
and implemented the new MEDYAN code, performed the simulations, analyzed results, and
wrote the manuscript. All authors contributed to the mean-field model, edited the manuscript,
and approved it for submission.
Competing interests
The authors declare no competing or financial interests.
27
Funding
This work was supported by the following grants from the National Science Foundation: NSF
1632976, NSF DMR-1506969, and NSF CHE-1800418.
Acknowledgments
The authors gratefully acknowledge A. Chandresekaran, Q. Ni, and H. Ni for their improvements
of the manuscript and helpful discussions.
References
[1] Daniel A Fletcher and R Dyche Mullins. Cell mechanics and the cytoskeleton. Nature,
463(7280):485, 2010.
[2] Isabelle Delon and Nicholas H Brown. Integrins and the actin cytoskeleton. Current opinion
in cell biology, 19(1):43 -- 50, 2007.
[3] S Burlacu, PA Janmey, and J Borejdo. Distribution of actin filament lengths measured
by fluorescence microscopy. American Journal of Physiology-Cell Physiology, 262(3):C569 --
C577, 1992.
[4] Marco Fritzsche, Christoph Erlenkamper, Emad Moeendarbary, Guillaume Charras, and
Karsten Kruse. Actin kinetics shapes cortical network structure and mechanics. Science
advances, 2(4):e1501337, 2016.
[5] Jonathan Stricker, Tobias Falzone, and Margaret L Gardel. Mechanics of the f-actin cy-
toskeleton. Journal of biomechanics, 43(1):9 -- 14, 2010.
[6] Daniel H Wachsstock, WH Schwartz, and Thomas D Pollard. Affinity of alpha-actinin for
actin determines the structure and mechanical properties of actin filament gels. Biophysical
journal, 65(1):205 -- 214, 1993.
[7] David Keller and Carlos Bustamante. The mechanochemistry of molecular motors. Bio-
physical Journal, 78(2):541 -- 556, 2000.
[8] Philipp J Albert, Thorsten Erdmann, and Ulrich S Schwarz. Stochastic dynamics and
mechanosensitivity of myosin ii minifilaments. New Journal of Physics, 16(9):093019, 2014.
[9] Jos´e Alvarado, Michael Sheinman, Abhinav Sharma, Fred C MacKintosh, and Gijsje H
Koenderink. Force percolation of contractile active gels. Soft matter, 13(34):5624 -- 5644,
2017.
[10] Mih´aly Kov´acs, Fei Wang, Aihua Hu, Yue Zhang, and James R Sellers. Functional di-
vergence of human cytoplasmic myosin ii kinetic characterization of the non-muscle iia
isoform. Journal of Biological Chemistry, 278(40):38132 -- 38140, 2003.
[11] Xuejun C Zhang and Wei Feng. Thermodynamic aspects of atp hydrolysis of actomyosin
complex. Biophysics reports, 2(5-6):87 -- 94, 2016.
[12] Thorsten Erdmann, Kathrin Bartelheimer, and Ulrich S Schwarz. Sensitivity of small
myosin ii ensembles from different isoforms to mechanical load and atp concentration.
Physical Review E, 94(5):052403, 2016.
[13] Martin McCullagh, Marissa G Saunders, and Gregory A Voth. Unraveling the mystery of
atp hydrolysis in actin filaments. Journal of the American Chemical Society, 136(37):13053 --
13058, 2014.
[14] Xin Li, Jan Kierfeld, and Reinhard Lipowsky. Actin polymerization and depolymerization
coupled to cooperative hydrolysis. Physical review letters, 103(4):048102, 2009.
28
[15] Sriram Ramaswamy. The mechanics and statistics of active matter. Annual Review of
Condensed Matter Physics, 1:323 -- 345, 2010.
[16] M Cristina Marchetti, Jean-Fran¸cois Joanny, Sriram Ramaswamy, Tanniemola B Liverpool,
Jacques Prost, Madan Rao, and R Aditi Simha. Hydrodynamics of soft active matter.
Reviews of Modern Physics, 85(3):1143, 2013.
[17] Terrell L Hill. Free energy transduction and biochemical cycle kinetics. Courier Corporation,
2004.
[18] Tzer Han Tan, Maya Malik Garbi, Enas Abu-Shah, Junang Li, Abhinav Sharma, Fred C
MacKintosh, Kinneret Keren, Christoph F Schmidt, and Nikta Fakhri. Self-organization of
stress patterns drives state transitions in actin cortices. arXiv preprint arXiv:1603.07600,
2016.
[19] Simone Kohler, Volker Schaller, and Andreas R Bausch. Structure formation in active
networks. Nature materials, 10(6):462, 2011.
[20] Volker Schaller, Christoph Weber, Christine Semmrich, Erwin Frey, and Andreas R Bausch.
Polar patterns of driven filaments. Nature, 467(7311):73, 2010.
[21] Jos´e Alvarado, Michael Sheinman, Abhinav Sharma, Fred C MacKintosh, and Gijsje H
Koenderink. Molecular motors robustly drive active gels to a critically connected state.
Nature Physics, 9(9):591, 2013.
[22] Ilya Prigogine. Introduction to thermodynamics of irreversible processes. New York: In-
terscience, 1967, 3rd ed., 1967.
[23] Ilya Prigogine and Gr´egoire Nicolis. Biological order, structure and instabilities. Quarterly
Reviews of Biophysics, 4(2-3):107 -- 148, 1971.
[24] Jeremy L England. Statistical physics of self-replication. The Journal of chemical physics,
139(12):09B623 1, 2013.
[25] Jeremy L England. Dissipative adaptation in driven self-assembly. Nature nanotechnology,
10(11):919, 2015.
[26] Nikolay Perunov, Robert A Marsland, and Jeremy L England. Statistical physics of adap-
tation. Physical Review X, 6(2):021036, 2016.
[27] Tal Kachman, Jeremy A Owen, and Jeremy L England. Self-organized resonance during
search of a diverse chemical space. Physical review letters, 119(3):038001, 2017.
[28] Jordan M Horowitz and Jeremy L England. Spontaneous fine-tuning to environment in
many-species chemical reaction networks. Proceedings of the National Academy of Sciences,
114(29):7565 -- 7570, 2017.
[29] Marco Baiesi and Christian Maes. Life efficiency does not always increase with the dissi-
pation rate. Journal of Physics Communications, 2(4):045017, 2018.
[30] Martial Balland, Alain Richert, and Fran¸cois Gallet. The dissipative contribution of myosin
ii in the cytoskeleton dynamics of myoblasts. European Biophysics Journal, 34(3):255 -- 261,
2005.
[31] Demosthene Mitrossilis, Jonathan Fouchard, Axel Guiroy, Nicolas Desprat, Nicolas Ro-
driguez, Ben Fabry, and Atef Asnacios. Single-cell response to stiffness exhibits muscle-like
behavior. Proceedings of the National Academy of Sciences, 106(43):18243 -- 18248, 2009.
[32] Brenton D Hoffman, Gladys Massiera, Kathleen M Van Citters, and John C Crocker. The
consensus mechanics of cultured mammalian cells. Proceedings of the National Academy of
Sciences, 103(27):10259 -- 10264, 2006.
[33] Laetitia Kurzawa, Benoit Vianay, Fabrice Senger, Timoth´ee Vignaud, Laurent Blanchoin,
and Manuel Th´ery. Dissipation of contractile forces: the missing piece in cell mechanics.
Molecular biology of the cell, 28(14):1825 -- 1832, 2017.
29
[34] Konstantin Popov, James Komianos, and Garegin A Papoian. Medyan: mechanochemical
simulations of contraction and polarity alignment in actomyosin networks. PLoS computa-
tional biology, 12(4):e1004877, 2016.
[35] Longhua Hu and Garegin A Papoian. Mechano-chemical feedbacks regulate actin mesh
growth in lamellipodial protrusions. Biophysical journal, 98(8):1375 -- 1384, 2010.
[36] James E Komianos and Garegin A Papoian. Stochastic ratcheting on a funneled energy
landscape is necessary for highly efficient contractility of actomyosin force dipoles. Physical
Review X, 8(2):021006, 2018.
[37] Giulio Ragazzon and Leonard J Prins. Energy consumption in chemical fuel-driven self-
assembly. Nature nanotechnology, page 1, 2018.
[38] Carlos Floyd, Garegin A. Papoian, and Christopher Jarzynski. A Discrete Approximation
to Gibbs Free Energy of Chemical Reactions is Needed for Accurately Calculating Entropy
Production in Mesoscopic Simulations. arXiv e-prints, page arXiv:1901.10520, January
2019.
[39] FJ Brooks and AE Carlsson. Nonequilibrium actin polymerization treated by a truncated
rate-equation method. Physical Review E, 79(3):031914, 2009.
[40] Carlos Floyd, Christopher Jarzynski, and Garegin Papoian. Low-dimensional manifold of
actin polymerization dynamics. New Journal of Physics, 19(12):125012, 2017.
[41] Christophe Leys, Christophe Ley, Olivier Klein, Philippe Bernard, and Laurent Licata.
Detecting outliers: Do not use standard deviation around the mean, use absolute deviation
around the median. Journal of Experimental Social Psychology, 49(4):764 -- 766, 2013.
[42] Peter J Huber. Robust statistics. Springer, 2011.
[43] Radek Erban, Mark B Flegg, and Garegin A Papoian. Multiscale stochastic reaction --
diffusion modeling: application to actin dynamics in filopodia. Bulletin of mathematical
biology, 76(4):799 -- 818, 2014.
[44] Longhua Hu and Garegin A Papoian. Molecular transport modulates the adaptive response
of branched actin networks to an external force. The Journal of Physical Chemistry B,
117(42):13388 -- 13396, 2013.
[45] Eric Jones, Travis Oliphant, and Pearu Peterson. {SciPy}: open source scientific tools for
{Python}. 2014.
[46] Eckhard Limpert, Werner A Stahel, and Markus Abbt. Log-normal distributions across
the sciences: Keys and clues: On the charms of statistics, and how mechanical models
resembling gambling machines offer a link to a handy way to characterize log-normal dis-
tributions, which can provide deeper insight into variability and probability -- normal or
log-normal: That is the question. AIBS Bulletin, 51(5):341 -- 352, 2001.
[47] Chikara Furusawa, Takao Suzuki, Akiko Kashiwagi, Tetsuya Yomo, and Kunihiko Kaneko.
Ubiquity of log-normal distributions in intra-cellular reaction dynamics. Biophysics, 1:25 --
31, 2005.
[48] Debasis Kundu and Anubhav Manglick. Discriminating between the log-normal and gamma
distributions. Journal of the Applied Statistical Sciences, 14:175 -- 187, 2005.
[49] Per Bak, Chao Tang, and Kurt Wiesenfeld. Self-organized criticality. Physical review A,
38(1):364, 1988.
[50] Per Bak and Chao Tang. Earthquakes as a self-organized critical phenomenon. Journal of
Geophysical Research: Solid Earth, 94(B11):15635 -- 15637, 1989.
[51] A Sornette and D Sornette. Self-organized criticality and earthquakes. EPL (Europhysics
Letters), 9(3):197, 1989.
30
[52] M Bottiglieri and C Godano. On-off intermittency in earthquake occurrence. Physical
Review E, 75(2):026101, 2007.
[53] Luca Cardamone, Alessandro Laio, Vincent Torre, Rajesh Shahapure, and Antonio DeSi-
mone. Cytoskeletal actin networks in motile cells are critically self-organized systems syn-
chronized by mechanical interactions. Proceedings of the National Academy of Sciences,
108(34):13978 -- 13983, 2011.
[54] Simon L Freedman, Glen M Hocky, Shiladitya Banerjee, and Aaron R Dinner. Nonequi-
librium phase diagrams for actomyosin networks. Soft matter, 14(37):7740 -- 7747, 2018.
[55] Daniel S Seara, Vikrant Yadav, Ian Linsmeier, A Pasha Tabatabai, Patrick W Oakes,
SM Ali Tabei, Shiladitya Banerjee, and Michael P Murrell. Entropy production rate is
maximized in non-contractile actomyosin. Nature communications, 9(1):4948, 2018.
[56] Christopher Battle, Chase P Broedersz, Nikta Fakhri, Veikko F Geyer, Jonathon Howard,
Christoph F Schmidt, and Fred C MacKintosh. Broken detailed balance at mesoscopic
scales in active biological systems. Science, 352(6285):604 -- 607, 2016.
[57] Tobias T Falzone, Savanna Blair, and Rae M Robertson-Anderson. Entangled f-actin
displays a unique crossover to microscale nonlinearity dominated by entanglement segment
dynamics. Soft matter, 11(22):4418 -- 4423, 2015.
[58] Ikuko Fujiwara, Dimitrios Vavylonis, and Thomas D Pollard. Polymerization kinetics of
adp-and adp-pi-actin determined by fluorescence microscopy. Proceedings of the National
Academy of Sciences, 104(21):8827 -- 8832, 2007.
[59] William M McFadden, Patrick M McCall, Margaret L Gardel, and Edwin M Munro. Fil-
ament turnover tunes both force generation and dissipation to control long-range flows in
a model actomyosin cortex. PLoS computational biology, 13(12):e1005811, 2017.
[60] Andrzej Duda and Adam Kowalski. Thermodynamics and kinetics of ring-opening poly-
merization. Handbook of ring-opening polymerization, pages 1 -- 51, 2009.
[61] Thorsten Erdmann, Philipp J Albert, and Ulrich S Schwarz. Stochastic dynamics of small
ensembles of non-processive molecular motors: The parallel cluster model. The Journal of
chemical physics, 139(17):11B604 1, 2013.
[62] Jonathon Howard. Mechanics of motor proteins and the cytoskeleton. Sinauer associates
Sunderland, MA, 2001.
[63] Takeshi Sakamoto, Martin R Webb, Eva Forgacs, Howard D White, and James R Sell-
ers. Direct observation of the mechanochemical coupling in myosin va during processive
movement. Nature, 455(7209):128, 2008.
[64] Paul A Dufort and Charles J Lumsden. How profilin/barbed-end synergy controls actin
polymerization: A kinetic model of the atp hydrolysis circuit. Cytoskeleton, 35(4):309 -- 330,
1996.
[65] Elena G Yarmola, Dmitri A Dranishnikov, and Michael R Bubb. Effect of profilin on actin
critical concentration: a theoretical analysis. Biophysical journal, 95(12):5544 -- 5573, 2008.
[66] Ewa Nowak and Roger S Goody. Kinetics of adenosine 5'-triphosphate and adenosine
5'-diphosphate interaction with g-actin. Biochemistry, 27(23):8613 -- 8617, 1988.
[67] HJ Kinosian, LA Selden, JE Estes, and LC Gershman. Nucleotide binding to actin. cation
dependence of nucleotide dissociation and exchange rates. Journal of Biological Chemistry,
268(12):8683 -- 8691, 1993.
[68] Lynn A Selden, Henry J Kinosian, James E Estes, and Lewis C Gershman. Impact of pro-
filin on actin-bound nucleotide exchange and actin polymerization dynamics. Biochemistry,
38(9):2769 -- 2778, 1999.
[69] Ron Milo and Rob Phillips. Cell biology by the numbers. Garland Science, 2015.
31
[70] G Nicolis and I Prigogine.
I (1977) self-organization in nonequilibrium systems. From
Dissipative structures to Order through Fluctuations, Mir, Moscow, Russia.
[71] Dilip Kondepudi and Ilya Prigogine. Modern thermodynamics: from heat engines to dissi-
pative structures. John Wiley & Sons, 2014.
[72] P Glansdorff and I Prigogine. Sur les propri´et´es diff´erentielles de la production d'entropie.
Physica, 20(7-12):773 -- 780, 1954.
[73] Charles S Peskin, Garrett M Odell, and George F Oster. Cellular motions and thermal
fluctuations: the brownian ratchet. Biophysical journal, 65(1):316 -- 324, 1993.
32
|
0908.3881 | 3 | 0908 | 2010-01-11T16:04:06 | Metastable Chimera States in Community-Structured Oscillator Networks | [
"physics.bio-ph",
"nlin.AO",
"q-bio.NC"
] | A system of symmetrically coupled identical oscillators with phase lag is presented, which is capable of generating a large repertoire of transient (metastable) "chimera" states in which synchronisation and desynchronisation co-exist. The oscillators are organised into communities, such that each oscillator is connected to all its peers in the same community and to a subset of the oscillators in other communities. Measures are introduced for quantifying metastability, the prevalence of chimera states, and the variety of such states a system generates. By simulation, it is shown that each of these measures is maximised when the phase lag of the model is close, but not equal, to pi/2. The relevance of the model to a number of fields is briefly discussed, with particular emphasis on brain dynamics. | physics.bio-ph | physics | Metastable Chimera States in Community-Structured
Oscillator Networks
Murray Shanahan
Department of Computing, Imperial College London, 180 Queen’s Gate, London SW7
2AZ, UK.
PACS areas: 05.45.Xt, 87.19.lm, 87.18.Sn, 89.75.Fb
Abstract
A system of symmetrically coupled identical oscillators with phase lag is presented,
which is capable of generating a large repertoire of transient (metastable) “chimera”
states in which synchronisation and desynchronisation co-exist. The oscillators are
organised into communities, such that each oscillator is connected to all its peers in the
same community and to a subset of the oscillators in other communities. Measures are
introduced for quantifying metastability, the prevalence of chimera states, and the
variety of such states a system generates. By simulation, it is shown that each of these
measures is maximised when the phase lag of the model is close, but not equal, to
" 2 .
The relevance of the model to a number of fields is briefly discussed, with particular
emphasis on brain dynamics.
!
1
Many complex systems, both natural and artificial, exhibit synchronisation
phenomena, which can be modeled using weakly coupled oscillators. Most previous
synchronisation studies focus on stability. Yet many complex systems (such as the
human brain) do not converge on stable synchronised states. Rather they are
metastable, temporarily dwelling in the vicinity of one stable state before
spontaneously migrating away from it towards another. A second feature of many
complex systems (including the brain) is competition. In the context of
synchronisation, this is manifest in so-called chimera states, where one coalition of
oscillators synchronises while rival coalitions of
identical oscillators are
desynchronised. This short paper presents the first model to exhibit both chimera
states and metastability. This is achieved by organising the oscillators into a
community structured
(echoing established brain
(or modular) network
connectivity findings). Measures are introduced to quantify the prevalence of
chimera states and metastability. These form the basis of an empirical study that
establishes the conditions in which metastability and chimera states are most
prevalent. The same study shows (using a novel measure) that the repertoire of
metastable states produced under these conditions is also maximised.
2
I. INTRODUCTION
Periodic phenomena involving the synchronisation of multiple variables are prevalent
both in nature and the human environment, and can be modeled mathematically as
systems of coupled oscillators [Pikovsky, et al. , 2001]. Among the rich variety of
behaviours such systems exhibit are states in which a set of identical symmetrically
coupled oscillators spontaneously partitions into one subset that is synchronised and
another subset that is desynchronised [Kuramoto & Battogtokh 2002; Abrams &
Strogatz, 2004; Abrams, et al., 2008]. A system that gives rise to these so-called
chimera states is a plausible model for a competitive process wherein a set of winners
forms an alliance to the exclusion of the rest of the population. Since competitive
processes of this sort dominate the dynamics of the brain, the economy, and the
ecosphere, they are of considerable scientific interest.
Typical studies of synchronisation in systems of coupled oscillators attempt to map their
various dynamical regimes and to pin down the conditions for entering those regimes
[Acebrón, et al., 2005]. Stable states, in which some or all of the oscillators are fully
synchronised, have attracted particular attention. However, in many complex systems,
extended periods of synchronisation are pathological. Prolonged synchronisation in the
brain, for example, is a symptom of seizure [Arthuis, et al., 2009]. This motivates the
study of metastability in systems of coupled oscillators [Niebur, et al., 1991; Bressler &
Kelso, 2001; Pluchino & Rapisarda, 2006; Kitzbichler, et al., 2009]. A system of
oscillators exhibits metastability if some or all of its members linger in the vicinity of a
synchronised state without falling into such a state permanently. Moreover, in a
3
complex milieu such as the brain, the economy, or the ecosphere, we should expect to
witness a large number of distinct metastable states.
The upshot of these considerations is that an adequate model of competitive periodic
phenomena in systems that are not frozen, static, or in seizure, should exhibit a non-
trivial repertoire of metastable chimera-like states. Although the topic has been
relatively neglected, metastability in oscillator networks has been described in the
literature before. For example, Niebur, et al. [1991] reported metastability in a large
network of weakly coupled oscillators. But their model included a noise term, and
metastability is promoted by the resulting thermal fluctuations [Kuramoto, 1984, Ch. 5].
More recently, Kitzbichler, et al. [2009], following the work of Kuramoto [1984],
characterised the metastability of a network of oscillators with critical coupling
strength. But the natural frequencies of the oscillators in their model are distributed
while models of chimera states deploy identical oscillators.
Reinforcing the well-estabished view that there is a crucial relationship between
connectivity and dynamics [Strogatz, 2001; Arenas, et al., 2008; Müller-Linow, et al.,
2008], the required properties are exhibited by the present (deterministic) model
because its oscillators (all identical) are arranged in a network with community structure
[Girvan & Newman, 2002]. That is to say, the nodes of the network (the individual
oscillators) are partitioned into subsets (communities or modules) whose members are
more densely connected with each other than with nodes outside their community. Since
both the functional and structural connectivity of the human brain are similarly modular
[Hagmann, et al., 2008], the system of oscillators presented here is a plausible model of
metastable neural synchronisation.
4
II. METHODS AND MEASURES
The present model comprises eight communities of 32 phase-lagged Kuramoto
oscillators with identical natural frequencies [Kuramoto, 1984; Acebrón, et al., 2005].
Each oscillator is fully connected to its own community, and has 32 random
connections to oscillators in other communities. Following the model of Abrams, et al.
[2008], the intra-community coupling strength is slightly higher than the inter-
community coupling strength. All connections are symmetrical.
The phase
"i of each oscillator i is governed by the equation
!
d"i
dt
= # +
1
N + 1
N
$
j =1
"j %"i %&)
K i, j sin(
where ω is the natural frequency of the oscillator, N is the total number of connections
!
K i, j is the coupling strength between oscillators i and j, and α is a fixed
per oscillator,
phase lag. In the present model, ω=1, N=63,
K i, j = u if i and j belong to the same
community,
K i, j = v if i and j belong to different communities and are connected, and
!
K i, j = 0 otherwise. Like Abrams, et al. [2008], we define two parameters A and β for
!
" = # / 2 $%. For the experiment
u + v = 1, and
A = u " v where
the model, such that
!
described here A was set at 0.2.
!
!
!
The level of synchrony within a community c at time t may be quantified according to
the measure
!
"c ( t ) = e
i#k ( t )
k $ c
!
5
where
!
"k ( t ) is the phase of oscillator k at time t and
k " c
over all k in c. This measure ranges from 0 to 1, where 0 is total desynchronisation and
1 is full synchronisation. (Note that
"c ( t ) quantifies instantaneous synchrony, and does
!
not provide information about coherence.)
denotes the average of f
f
!
"c ( t ) for all the communities at discrete intervals, it is possible to quantify
By sampling
both the level of metastability in the system and the prevalence of chimera-like states
(Fig. 1). Let C be the set of all M communities, and assume
"c ( t ) is sampled at times
!
t " {1...T} for each
"met (c )
c " C . If we fix the community c and estimate the variance
of
"c ( t ) over all time points
t " {1...T} , we get an indication of how much the
!
synchrony in c varies over time. The average of this variance estimate over the set C of
!
!
all communities is an index of the metastability of the overall system (denoted λ). So
!
we have
!
!
" = #met C
where
!
"met (c ) =
1
T # 1
&
t %T
($c ( t ) # $c T
) 2 .
"c ( t ) over all
"chi ( t ) of
Conversely, if we fix the time t and estimate the variance
!
communities in C, we get an instantaneous indication of how chimera-like the system is
at time t. The average of this variance estimate is an index of how chimera-like a typical
!
!
state of the system is (denoted χ). So we have
" = #chi T
!
6
where
"chi ( t ) =
1
M # 1
&
c %C
($c ( t ) # $( t )
) 2 .
C
Note that if a population c of oscllators is either completely synchronised or completely
!
desynchronised then
"met (c ) = 0 . If c spends equal time in all stages of synchronisation
then it presents a uniformly distributed
"met (c )
12 # 0.083 . A value of
"c with
"met (c ) = 1
greater than
12 is possible if
"c has a multi-modal distribution with high peaks at both
!
1
shoulders. But assuming such distributions do not arise, and as long as the population
!
!
!
size is large, we may regard a uniform distribution as indicative of maximum
!
!
12 # 0.083 . (The population size must be large, since, for
metastability, yielding
"max = 1
example, a pair of decoupled (free-running) oscillators with different frequencies is not
metastable yet has uniformly distributed
"c . By contrast, episodes of high synchrony
!
are rare in a large population of decoupled oscillators with different frequencies,
ensuring a non-uniformly distributed
"c . In the present model, the community size is
!
large, and of course the oscillators are not decoupled.)
!
"chi . If at some time t all of the communities in the
Similar considerations apply to
system are either fully synchronised or fully desynchronised then
"chi ( t ) = 0 . A
“perfect” chimera state might be characterised as one in which exactly half the
!
communities are fully synchronised and half are fully desynchronised (as in the model
!
of Abrams, et al. (2008)), which will yield
7 # 0.2857 . Although high values
"chi ( t ) = 2
of
"chi can be obtained in the present model, due to its metastability they are only
" averages
transient. So because
"chi over time, we should not expect it to approach
!
" if
this maximum. Instead, a metastable system can be considered to attain maximum
!
!
7
!
!
it spends half its time in a maximally chimera-like state and half its time in a minimally
chimera-like state, which yields
7 # 0.1429 .
"max = 1
In addition to assessing synchronisation within a community, it is possible to quantify
!
pairwise synchronisation across communities. In particular, for every pair of
communities a and b, we can consider
(
"a ,b ( t ) = 1
2 e
i#k ( t )
k $ a
i#k ( t )
+ e
) .
k $ b
"b ( t ) are
"a ( t ) and
"a ,b ( t ) , which ranges from 0 to 1, will only be high if
Note that
!
individually high and communities a and b are synchronised with each other. If a and b
are fully synchronised internally but 180° out of phase with each other then
"a ,b ( t ) = 0 .
!
!
!
Although λ and χ taken together can detect the occurrence of metastable chimera states,
!
neither measure can distinguish a system that repeatedly visits the same metastable
chimera state from a system that has a large repertoire of metastable chimera states. One
way to quantify this repertoire is to assess how “mixed up” is the set of coalitions a
system produces over a period of time. Accordingly, we define the (normalised)
H C of a system by the equation
coalition entropy
H C = "
!
1
log 2 S
$
s# S
p( s) log 2
( p( s))
p( s) is the
where S is the set of distinct coalitions the system can generate and
!
probability of coalition s arising in any given time point. The measure is normalised to
lie to between 0 and 1. In the context of the present model we can consider coalitions of
!
synchronised communities. A coalition s is said to arise at time t if
"c ( t ) > # for all
8
!
(a)
(b)
FIG. 1: The behaviour of the model for randomly generated values of β between 0
" 4 over 500 trials. Initial phases were randomised for each trial. Both
and
metastability index (λ) and chimera index (χ) are close to zero for
" = 0 , peak at
" = 0.1, and tail off rapidly.
around
!
!
c " s , where γ is a synchronisation threshold (we shall use
" = 0.8 ). Clearly for a
!
!
M possible coalitions, so
system of M communities there are
M
log 2 S = M . If all
2
2
possible coalitions arise with equal probability, we have
H C = 1. On the other hand, if
!
the system resides permanently in the same state (whether fully synchronised, fully
!
!
!
desynchronised, or any sort of chimera-like state), then only one coalition arises and we
!
H C = 0 .
have
III. RESULTS
!
A series of 1000-step trials of for a range of values of β was carried out. 500 trials were
performed for randomly generated values of β ranging from 0 to
" 4 . All numerical
simulation was carried out using the 4th-order Runge-Kutta method with a step size of
!
9
FIG. 2: Global synchrony (Φ) and coalition entropy for the set of trials depicted in
Figure 1. Coalition entropy presents a similar profile to metatsability and chimera
index (Fig. 1), although it peaks slightly later, at around
" = 0.15 .
0.05. Intra-community coupling assignments and initial phases were randomised for
!
each trial. The internal synchrony
"c ( t ) was calculated for each community at 5-step
H C . In addition, an index
intervals, and the resulting data was used to compute λ , χ and
of global synchrony Ψ was calculated, taken as the average of
"c ( t ) over all times and
!
communities.
!
!
The results are presented in scatter plots of Figs. 1 & 2. These figures suggest that the
model behaves as advertised, and is capable of generating a large repertoire of
metastable chimera-like states, but only when β falls within a certain narrow range.
Metastability and chimera indices are maximised when
0.05 < " < 0.15 (Fig. 1), at
" 2 ), the
" = 0 (ie: the phase lag is exactly
0.6 < " < 0.7 (Fig. 2). When
which point
system finds it hard to synchronise at all, and each of the measures is correspondingly
!
low. At the other end of the scale, when
" > # 8 , the system tends towards full
!
!
!
synchronisation in all communities, and metastability, chimera index, and coalition
entropy all tail off accordingly. Coalition entropy peaks slightly later than the other two
!
0.1 < " < 0.2 (Fig. 2).
measures, with
!
10
As with other systems of oscillators that exhibit metastability, fluctuations of
synchronisation and desynchronisation are prevalent only in a narrow, critical region
that resembles a thermodynamic phase transition from order to disorder [Kuramoto,
1984, Ch. 5]. The area to the left of the critical region is the disordered regime, wherein
the phases of the oscillators are subject to predominantly repelling forces, while the area
to the right of the critical region is the ordered regime, wherein attracting forces
dominate. In the critical region there is a balance of repelling and attracting forces. But
this balance is not static, and the dominant force applicable to each community
alternates between phase attraction and phase repulsion. Since the oscillators have
identical natural frequencies and there is no external stochastic perturbation, the reasons
for this are not clear, and further work is required to understand the mechanisms
underlying the model’s behaviour.
To gain some preliminary insight into the behaviour of the model in the critical region,
we shall examine a single run in detail. Figure 3 (a) shows the evolution of synchrony
") within all eight oscillator communities over a 200 time-step interval in a typical
(
trial with
" # 0.1. The trial is representative of the dynamical regime of most interest to
HC = 0.6341, and
" = 0.0542 ,
" = 0.0525 ,
us here. The statistics for this trial are:
" = 0.5798 . The system exhibits several chimera-like states in which some oscillator
!
communities are highly synchronised while others are desynchronised. From time 250
!
!
!
to 270, for example three communities are highly synchronised (φ > 0.8) and three are
desynchronised (φ < 0.4), while two communites have intermediate values of φ. A
similarly clear chimera-like state can be seen from time 310 to 330, but with a different
combination of synchronised communities.
!
!
11
(a)
(b)
" # 0.1. (a) Intra-community
FIG. 3: A 200 time-step period in a typical run with
synchrony: The system exhibits several clear chimera-like states. From time 250 to
270, for example three communities are highly synchronised and three are
desynchronised. (b) Inter-community synchrony: Pairwise synchrony is plotted
between one selected community (shown in black) and each of the eight
!
communities (including itself). From time 310 to 330 the selected community is
synchronised with several others, forming a temporary coalition.
Fig. 4 shows how synchrony is distributed over time for each of the eight communities
in the same sample run. As expected from Fig. 3 (a), none of the communities spends
the majority of its time in any one stage of internal synchronisation. Despite a tendency
towards synchrony (a rightwards skew), the resulting distributions all have high
variance. Its noteworthy, however, that different communities exhibit different profiles,
12
FIG. 4: The distribution of
" for each community in the run depicted in Figure 3.
The distributions all have variances far from zero, indicating metastability in the
sense that the oscillator communities spend time in all stages of synchronisation.
!
with Community 2, for example, showing a more pronounced tendency to internally
synchronise than its peers.
The plots in Fig. 3 (a) are highly irregular, and there is no clear pattern to the
coincidence of peaks and troughs, which suggests that the system is capable of
generating a large repertoire of coalitions of simultaneously synchronised communities.
By contrast, the equivalent plot for a “breathing” chimera state has a regular rhythm
(see Fig. 2 of [Abrams, et al., 2008]). The lack of any discernible regularity in the
sequence of metastable states visited by the system hints at both chaotic itinerancy
[Kaneko & Tsuda, 2003; Tsuda, et al., 2004] and dynamical complexity [Tononi, et al.,
1998; Seth, et al., 2006; Shanahan, 2008b], but further work would be required to
establish these properties rigorously.
13
Further insight into the generation of coalitions by examining the ebb and flow of
pairwise synchronisation (
"). Fig 3 (b) shows, for the same 200-step period of the same
trial, the pairwise synchrony
"a ,b between a selected community a and all communities
b (including a itself). As noted in the previous section, the pairwise synchronisation
!
between two oscillator communities will be low if either community has low internal
!
synchronisation. In the troughs near times 260 and 340, all pairwise synchrony
measures are low because the internal synchrony of community a is itself low. At other
times, the internal synchrony of a is high, as is the pairwise synchrony between a and
various peers. Sets of such synchronised peers constitute temporary coalitions, whose
constitution varies over time. At time 290, for example, a is in a coalition alongside five
of its peers, with two communities excluded. But by time 320 three communities have
been expelled from the coalition, while the two previous exclusions have been recruited
into its membership.
IV. DISCUSSION
Although the essential characteristic of the model — the ability to generate a large
repertoire of metastable chimera states — reflects properties common to many real-
world complex dynamical systems, the task remains of mapping each of those systems
onto the model. For example, it has been proposed that synchronised oscillations in the
brain permit effective co-operation among distinct populations of neurons [Fries, 2005;
2009; Womelsdorf, et al., 2007], while phenomena such as binocular rivalry,
inattentional blindness, and the Stroop effect attest to their competitive character. In
other words, the dynamics of the brain seems to arise from the interplay of co-operation
and competition, resulting in the formation of synchronised coalitions [Doesburg, et al.,
14
2009]. Moreover, thanks to an animal’s ongoing activity, its brain is endlessly subject to
an open-ended variety of perturbations, and to respond effectively these coalitions must
be in constant flux. The dynamics of the present model exhibits the same combination
of features. However, to date there is no neurologically detailed model to match,
although several existing spiking neuron models deal with relevant issues, such as
cortical competition [Dehaene, et al., 2003; Deco & Rolls, 2005; Shanahan, 2008a],
community structure [Shanahan, 2008b], and the interplay of synchronisation and
desynchronisation [Tsuda, et al., 2004].
The findings reported here raise a number of questions to be addressed in future work.
The issues of chaotic itinerancy and dynamical complexity have already been
mentioned. Another urgent task is to develop a theoretical understanding of the
phenomena described. The obvious starting point is the analytical treatment of Abrams,
et al. [2008], whose model is the basis for the present work. However, analogous results
may be difficult to obtain for the present model, given its more complex network
structure and the variety of synchronisation effects it displays. Finally, it would be
fruitful to investigate the occurrence of metastable chimera states in networks with
hierarchical community structure (especially as brain networks exhibit this property
[Zhou, et al., 2006; Ferrarini, et al, 2009]), and to attempt to characterise the “path” to
such states in the manner of Arenas, et al. [2006].
REFERENCES
1. D. M. Abrams, R. Mirollo, S. H. Strogatz, and D. A. Wiley, Phys. Rev. Lett. 101,
084103 (2008).
15
2. D. M. Abrams and S. H. Strogatz, Phys. Rev. Lett. 93, 174102 (2004).
3. J. A. Acebrón, L. L. Bonilla, C. J. P. Vicente, F. Ritort, and R. Spigler, Rev. Modern
Phys. 77, 137 (2005).
4. A. Arenas, A. Díaz-Guilera, J. Kurths, Y. Moreno, and C. Zhao, Phys. Reports 469,
93 (2008).
5. A. Arenas, A. Díaz-Guilera, and C. J. Pérez-Vicente, Phys. Rev. Lett. 96, 114102
(2006).
6. M. Arthuis, L. Valton, J. Régis, P. Chauvel, F. Wendling, L. Naccache, C. Bernard,
and F. Bartolomei, Brain 132, 2091 (2009).
7. S. L. Bressler and J. A. S. Kelso, Trends Cogn. Sci. 5 (1), 26 (2001).
8. G. Deco and E. T. Rolls, J. Neurophysiol. 94, 295 (2005).
9. S. Dehaene, C. Sergent, and J.-P. Changeux, Proc. Nat. Academy Sci. 100, 8520
(2003).
10. S. M. Doesburg, J. J. Green, J. J. McDonald and L. M. Ward, PLoS ONE 4, e6142
(2009).
11. L. Ferrarini, I. M. Veer, E. Baerends, M.-J. van Tol, R. J. Renken, N. J. A. van der
Wee, D. J. Veltman, A. Aleman, F. G. Zitman, B. W. J. H. Penninx, M. A. van
Buchem, J. H. C. Reiber, S. A. R. B. Rombouts, and J. Milles, Hum. Brain. Mapp.
30, 2220 (2009).
12. P. Fries, Trends Cogn. Sci. 9 (10), 474 (2005).
16
13. P. Fries, Ann. Rev. Neurosci. 32, 209 (2009).
14. M. Girvan and M. E. J. Newman, Proc. Natl. Acad. Sci. U.S.A. 99, 7821 (2002).
15. P. Hagmann, L. Cammoun, X. Gigandet, R. Meuli, C. J., Honey, V. J. Wedeen, and
O. Sporns, PLoS Biol. 6, e159 (2008).
16. K. Kaneko and I. Tsuda, Chaos 13, 926 (2003).
17. M.G.Kitzbichler, M.L.Smith, S.R.Christensen, and E.Bullmore, PLoS Comp. Biol.
5, e1000314 (2009).
18. Y. Kuramoto, Chemical Oscillations, Waves, and Turbulence, Springer-Verlag
(1984).
19. Y. Kuramoto and D. Battogtokh, Nonlinear Phenom. Complex Syst. 5, 380 (2002).
20. M. Müller-Linow, C. C. Hilgetag, and M.-T. Hütt, PLoS Comp. Biol. 4, e1000190
(2008).
21. E. Niebur, H. G. Schuster, and D. M. Kammen, Phys. Rev. Lett. 67, 2753 (1991).
22. A. Pikovsky, M. Rosenblum, and J. Kurths, Synchroniztion: A Universal Concept in
Nonlinear Sciences, Cambridge University Press (2001).
23. A. Pluchino and A. Rapisarda, Physica A 365, 184 (2006).
24. A. K. Seth, E. M. Izhikevich, G. N. Reeke, and G. M. Edelman, Proc. Natl. Acad.
Sci. U.S.A. 103, 10799 (2006).
25. M. Shanahan, Conscious. Cogn. 17, 288 (2008).
17
26. M. Shanahan, Phys. Rev. E 78, 041924 (2008).
27. S. H. Strogatz, Nature 410, 268 (2001).
28. G. Tononi, G. M. Edelman, and O. Sporns, Trends Cogn. Sci. 2, 474 (1998).
29. I. Tsuda, H. Fuji, S. Tadokoro, T. Yasuoka, and Y. Yamaguti, J. Integr. Neurosci. 3,
159 (2004).
30. T. Womelsdorf, J.-M. Schoffelen, R. Oostenveld, W. Singer, R. Desimone, A. K.
Engel, and P. Fries, Science 316, 1609 (2007).
31. C. Zhou, L. Zemanová, G. Zamora, C. C. Hilgetag, and J. Kurths, Phys. Rev. Lett.
97, 238103 (2006).
18
|
1912.03495 | 1 | 1912 | 2019-12-07T12:51:32 | Randomness and optimality in enhanced DNA ligation with crowding effects | [
"physics.bio-ph"
] | Enzymatic ligation is essential for the synthesis of long DNA. However, the number of ligated products exponentially decays as the DNA synthesis proceeds in a random manner. The controlling of ligation randomness is of importance to suppress exponential decay and demonstrate an efficient synthesis of long DNA. Here, we report the analysis of randomness in sequential DNA ligations, named qPCR-based STatistical Analysis of Randomness (qPCR-bSTAR), by a probability distribution of ligated DNA concentration. We show that the exponential decay is suppressed in a solution of another polymer and DNA ligation is activated at an optimal crowded condition. Theoretical model of kinetic ligation explains that intermolecular attraction due to molecular crowding can be involved in the optimal balance of the ligation speed and the available ligase. Our finding indicates that crowding effects enhance the synthesis of long DNA that retains large genetic information. | physics.bio-ph | physics |
Randomness and optimality in enhanced DNA ligation with crowding effects
Takaharu Y. Shiraki1, Ken-ichiro Kamei2, and Yusuke T. Maeda1
1Kyushu University, Department of Physics, Motooka 744, Fukuoka 812-0395, Japan and
2Kyoto University, Institute for Integrated Cell-Material Sciences (iCeMS),
Yoshida-Ushinomiyacho, Sakyo-ku, Kyoto 606-8501, Japan
(Dated: December 10, 2019)
Enzymatic ligation is essential for the synthesis of long DNA. However, the number of ligated
products exponentially decays as the DNA synthesis proceeds in a random manner. The controlling
of ligation randomness is of importance to suppress exponential decay and demonstrate an efficient
synthesis of long DNA. Here, we report the analysis of randomness in sequential DNA ligations,
named qPCR-based STatistical Analysis of Randomness (qPCR-bSTAR), by a probability distribu-
tion of ligated DNA concentration. We show that the exponential decay is suppressed in a solution
of another polymer and DNA ligation is activated at an optimal crowded condition. Theoretical
model of kinetic ligation explains that intermolecular attraction due to molecular crowding can be
involved in the optimal balance of the ligation speed and the available ligase. Our finding indicates
that crowding effects enhance the synthesis of long DNA that retains large genetic information.
INTRODUCTION
Genetic polymers such as DNA or RNA need to be
elongated in order to store information in their sequences.
To make a long DNA polymer, sequential ligation of DNA
fragment catalyzed by a ligase enzyme is an essential step
for the repairing DNA break of genome in living cells
and the in vitro synthesis of the artificial genome[1][2][3].
As end-to-end ligation occurs in a test tube is random
process, new short fragments are connected from both
ends to make longer DNA but the concentration of lig-
ated polymers decays with the number of newly added
monomers [4][5]. Because ligation randomness restricts
the abundance of the polymerized product, it has been
long discussed how long genetic polymers are synthesized
in nature and technology through random enzymatic lig-
ation.
One scenario to drive an efficient synthesis of long
DNA is the local increase of short DNA fragments by
using physical transport effects. It has been studied that
DNA fragments whose size range from a few tens to a few
hundreds of base pairs are accumulated due to molecu-
lar transport and thermal convection under temperature
gradients[6][7][8][9]. A local temperature gradient or so-
lute concentration gradient induces the directed motion
of DNA as a solute and in turn the trapped DNA shows
an exponential increase in the concentration[10][11]. This
DNA trapping is thought to compensate the significant
decrease of long DNA polymers.
The other relevant scenario is the effect from the
coexisted polymer, known as molecular crowding[12].
When DNA coexist with large concentrations of poly-
mers with a smaller gyration radius, attractive inter-
actions occur among DNA due to an excluded vol-
ume effect[13][14]. This molecular crowding has a wide
range of effects from structural changes and catalytic ac-
tivity of protein enzymes[15][16][17][18] to subdiffusive
motion[19][20]. The interplay between molecular crowd-
ing and the DNA ligation is of importance, but the lack
of experimental methods to measure and control the
randomness of DNA ligation is major bottleneck to re-
veal the physical mechanism for long DNA synthesis in
crowded condition.
In this study, we investigate the effect of molecular
crowding on the long DNA synthesis by using end-to-end
ligation of short DNA fragments with coexisting polymer
(polyethylene glycol, PEG). To reveal the randomness in
ligation, the concentration of ligated DNA products is
measured by a quantitative PCR (qPCR)-based method.
This method quantifies the concentration distribution of
ligated DNA products with the number of ligated joints
per molecule. We find that the ligated DNA fragments
show the gradual exponential decay in the presence of
crowding agents, as the number of ligated joints increases
and the production of long DNA is much enhanced with
smaller decay rate. We further test the optimal condi-
tion of molecular crowding by changing the PEG con-
centration and find that the exponential decay becomes
larger and the ligation efficiency is decreased due to the
reduction of freely available ligase as PEG concentration
is larger than 10%. The control of such ligation ran-
domness may be useful not only for the synthesis of ar-
tificial genomes in vitro but also the understanding of
enzymatic reactions in crowded conditions such as intra-
cellular space.
RESULTS
The analysis of randomness in enzymatic ligation
We examine the enzymatic ligation of short DNA
fragments under macromolecular crowding. To analyze
how the long DNA strands are efficiently synthesized in
crowding condition, an analytical method that can an-
alyze a ligated DNA product on its length and concen-
tration is needed. This technical requirement motivates
us to develop new method for probing the probability
distribution of DNA concentration with respect to size,
named as the qPCR-based Statistical Analysis of Ran-
2
left in Fig. 1(a)). DNAs with sticky ends were lig-
ated by Taq DNA ligase. For instance, B fragment
and C fragment became a single DNA chain BtC
after ligation since bc and bc formed a hybridized
joint.
• Third step: Afterwards, the enzyme ligated DNA
fragments in an alphabetical order (bottom right in
Fig. 1(a))). Ligation of 11 short DNA fragments
produced ligated DNA molecules with 66 different
sequences and lengths. The ordered sequence was
represented by a simple description as XtY, mean-
ing that short DNA fragments were connected from
the X fragment to the Y fragment. For instance, a
DNA that includes A-B-C-D sequence is referred
to as AtD. While K-K refers to the K fragment
alone. In the following description, DNA length is
the number of short DNA fragments contained per
molecule.
• Fourth step: In order to find the probability distri-
bution of the concentration of ligated DNA in its
length, DNA fragments having the K sequence at
the end was extracted and their concentration was
measured by qPCR (Bottom left in Fig. 1(a))).
We modified the blunt end of K fragment with bi-
otin and extracted 11 target DNA by the pull-down
method, using strong binding to streptavidin. The
obtained DNA-magnetic beads were used for qPCR
analysis.
The concentration of selected DNA products containing
K fragment reports the distribution function of ligated
DNA concentration. The number of unit DNA fragments
L (1 ≤ L ≤ 11) is proportional to the number of ligated
joints per molecule (that is L − 1). The concentration
of DNA obtained by qPCR analysis C(L) is the cumula-
n=L cDN A(n). Normalized cumula-
tive distribution function (CDF) of the ligated products
F (L) = C(L)/C(1) was used to test whether F (L) de-
cays exponentially with respect to DNA size. CDF F (L)
gives an experimental clue about the randomness behind
enzymatic ligation (Fig. 1(b)).
tive concentration P11
FIG. 1. Principle of qPCR-bSTAR method. (a) Experimental
procedure for qPCR-bSTAR. We prepared short DNA frag-
ments (step 1) and performed DNA ligation (step 2). We ob-
tained the 66 species of ligated products (step 3) and then se-
lected subpopulation of 11 DNA species by pull-down method.
M.B. stands for magnetic beads. (b) qPCR-bSTAR quanti-
fies the concentration of ligated DNA products as a cumula-
tive distribution function (CDF) of DNA concentration. By
using primers that detect sequences, including jk(red box),
we measured the sum of concentrations of JtK, · · · , BtK
and AtK. (c) Experimentally measured CDF as a function
of DNA length L after ligation in a buffer solution. The
experimental data was fitted well with exponential function
exp(−0.84 · L). Error bars represent the standard deviation
from three independent experiments.
domness (qPCR-bSTAR).
The experimental protocol of qPCR-bSTAR is de-
scribed in Fig. 1(a).
• First step: Short DNA fragments are synthesized
1(a)).
by PCR amplification (Top left in Fig.
These DNA fragments had 11 different sequences
and were distinguished by labeling from A, B, · · · ,
K. Their base pair (bp) number were all the same
size and are 258 bp. To make ligated joint be-
tween DNA fragments, their both ends had over-
hang structure called sticky ends. Recognition site
by restriction enzyme shown in letters, ab or bc · · ·
(ab is complementary sequence of ab), was added
to their ends. Sticky ends were obtained as DNA
substrate for subsequent end-to-end ligation.
• Second step: DNA ligation was performed (Bottom
Random DNA ligation with crowding effect
We analyzed the ligation reaction in a buffer solution
by using qPCR-bSTAR method. CDF F (L) was plotted
for the DNA size L (Fig. 1(c)) and it showed an expo-
nential decay, F (L) = exp(−αL). The exponential decay
of F (L) reflects that the ligation is a random reaction not
depending on DNA length itself. A similar exponential
decay was also observed in capillary tube electrophore-
sis analysis though we could detect DNA fragments only
shorter than L ≤ 6 (Fig. S1)[23].
Next question is how this exponential decay of F (L)
can be suppressed due to physically relevant effects. It
has long been studied that molecular crowding of inert
3
FIG. 2. The analysis of DNA ligation in polymer solutions by using qPCR-bSTAR method. (a) Schematic illustration of DNA
ligation in a polymer solution with crowding agent. The green particle represents a crowding agent. We used polyethylene
glycol (PEG) as a crowding agent. (b) Experimental result of CDF F (L) of ligated DNA in various PEG concentrations. The
color code represents the concentration of crowding agent PEG, from 0% (red), 7.5% (green), 10.0% (blue) to 15.0% (purple).
The fitting curves is F (L) = exp(−αL) with α = 0.84 in 0% PEG, α = 0.63 in 7.5% PEG, α = 0.29 in 10.0% PEG, α = 0.99
in 15.0% PEG. Error bars represent the standard deviation from three independent experiments. (c) Boxplot of mean DNA
length hLi in PEG concentration. Statistical analysis of Mann-WhitneyU-test was performed.
polymers increases enzyme activity[12], but it is not clear
whether ligation reactions occur randomly even in such a
crowded environment (Fig. 2(a)). To clarify this point,
we performed qPCR-bSTAR assay for the DNA ligation
with various concentration of polyethylene glycol 6000
(PEG) as a crowding agent from 2.5%to 15.0% (w/w).
As shown in Fig. 2(b), F (L) of the ligated DNA also
showed the exponential decay from 2.5% to 10.0% PEG,
but the slope of exponential decay was more gradual than
the decay rate in a buffer solution(0% PEG). This result
means that the ligation reaction occurred at random even
in the crowded environment but the number of reaction
products was much increased. Moreover, as PEG concen-
tration reached 15.0%, F (L) showed turned to stretched
exponential decay where the tail of the distribution func-
tion was stretched slightly (Fig. 2(b), purple). It sug-
gests that DNA ligation has optimal balance of enzymatic
reaction and molecular crowding, around at 10.0% PEG,
for efficient synthesis of the longest product.
According to the exponential distribution function
F (L) in various PEG concentrations, the mean DNA
length (the average number of ligations per molecule) hLi
was analyzed. Fig. 2(c) shows that hLi was increased by
the addition of PEG and had a peak at 10.0% PEG,
but it sharply decreased at higher concentrations. This
optimal concentration of PEG implies that the molecu-
lar crowding enhances the synthesis of longer DNA but
the effect underlying this enhanced ligation was gradually
suppressed at the over-crowded environment.
Kinetics of DNA ligation
The qPCR-bSTAR analysis allows us to find that the
DNA ligation is random both in the buffer solution and
the crowded environments, but the mean length hLi can
be affected by the concentration of PEG. However, why
is there optimal concentration of PEG at enhanced DNA
ligation? To reveal its underlying mechanism, we mea-
sured the kinetic increase of ligated products with two
pairs of K fragment and another DNA that can be jointed
to K fragment.
We firstly chose the J fragment of same length 258 bp
as the short DNA substrate. The rate of the enzymatic
reaction in the presence of coexisting PEG was evalu-
ated by measuring the ligated product of the K and J
fragments. The ligation reaction started in an aqueous
solution (0% PEG), and a small fraction of solution was
taken every 1 min and the concentration of the prod-
uct JtK was measured by qPCR (Fig. 3(a)). The con-
centration of JtK increased linearly with time, and the
slope of reaction kinetics increased as the concentration
of J (Fig. 3(a)). Such linear increase with time was also
observed even when PEG 10% was added (Fig. 3(b)).
However, when the PEG concentration exceeded 15%, al-
though the product JtK increases linearly over time, the
slope of reaction curve became independent of the initial
concentration of J (Fig. 3(c)). The ligation proceeded at
a constant rate regardless of the DNA concentration at
15% PEG, indicating that ligation reaction is no longer
dependent on the number of substrate.
Next, we chose the AtJ fragment that is ligated all 10
fragments from A to J, which is 10 times longer than
either the K and J fragment. By analyzing the liga-
tion kinetics of the K and AtJ, we found the linear in-
crease of ligation product AtK in both 0% (Fig. 3(d))
and 10.0% PEG (Fig. 3(e)), and the slope of kinetics
was increased with the initial amount of AtJ fragment.
Interestingly, in 15.0% PEG, the slope of kinetics was
suppressed as the initial amount of AtJ fragment in-
creased (Fig. 3(f)). These results indicate that the pres-
ence of macromolecules confers a size dependence of the
substrate molecules in the DNA ligation reaction (Figs.
3(g)-(i)).
Kinetic model of DNA ligation with crowding effect
We propose theoretical model where two effects are
taken into account to explain the concentration depen-
dence of the production rate v and the DNA concentra-
tions of J and AtJ. Michaelis-Menten model is known
to explain the kinetics of enzymatic reaction by assum-
ing enzyme-substrate complex as an intermediate prod-
uct. We extended this model by considering intermediate
products in order to understand the optimal condition for
enhanced DNA ligation in crowding conditions.
We define [Si] as the concentration of DNA i where i
stands for one arbitrary sequence from all 66 DNA se-
quences. The length of DNA i is Li. Suppose that liga-
tion reaction mainly consists of three chemical processes:
the First process is the hybridization of complementary
DNA pair, for instance, J and K fragments. The second
step is the complex formation of DNA pair and the ligase
enzyme and the third is to trigger the end-to-end DNA
ligation at the DNA-ligase complex. To explain the ki-
netics v ∝ [SX ]−1 as seen in Figs. 3(h) and 3(i), we need
to take into account additional pathways to the above
three ordinary steps.
One additional effect newly considered is the forma-
tion of partial hybridization of a pair of DNA where the
sequences of sticky ends are not completely complemen-
tary (Fig. 4(a), Effect 1). The sticky ends have 4 base
for hybridization, but we assume that the DNA strands
can make partially hybridized pairing even in there is
one or two base pair mismatch. Although the partial
hybridization is weaker in a buffer solution, molecular
crowding due to coexisting polymer stabilizes the partial
hybridization.
Another pathway newly considered is that the non-
specific interaction of DNA and the ligase. A ligase
finds hybridized sticky ends and makes new join between
4
two DNA fragments. We assume that the ligase protein
is trapped onto DNA fragment in sequence-independent
manner due to the molecular crowding. In dense PEG so-
lution, molecular crowding induces the attractive interac-
tion between the ligase (known as depletion force[13][14]).
The longer DNA has more space to attract ligase pro-
teins and traps the active enzyme away from the sticky
ends (Fig. 4(a), Effect 2). This nonspecific binding onto
DNA reduces the freely available enzyme in a solution,
meaning that the effective concentration of ligase is de-
creased in the presence of long DNA in dense PEG so-
lution. These pathways do not contribute to make lig-
ated products but change the reaction diagram, at end-
to-end ligation. By taking into account these pathways
appeared in crowded environment, the speed of ligation
reaction with two DNA fragments (Si and Sj) is given
by,
vij =
k+2[E0][Si][Sj]
(1)
Ψ(cpeg)Θ(S, cpeg; L) +P(l,m)∈P [Sl][Sm]
where S represents the concentrations of all 66 DNA
([S1], [S2], · · ·
[S66]), [E0] is the total concentration of
ligase, cpeg is the concentration of PEG. L represents
the lengths of all 66 DNA (L1, L2, · · · , L66). P is a set of
all possible DNA pair whose sequence of sticky ends are
complementary. The reaction constant of ligation k+2 de-
creases exponentially as a function of cpeg[23][3]. Ψ(cpeg)
is the dissociation constant among freely available lig-
ase [E], [Si] and [Sj], and Θ(S, cpeg; L) represents the
ratio of ligase trapped on DNA and the freely available
ligase[23].
JK with J fragment (and v−1
We performed the numerical calculation of vij at var-
ious PEG concentration cpeg by changing the concentra-
tion of [Si] while keeping [Sj]. Experimental results of
Figs. 3(g) and 3(h) proposes two cases of crowded lig-
ation as model systems: First one is the ligation of two
short DNA fragments i and j that correspond to J and
K fragments, and second one is the ligation of long DNA
and short DNA that correspond toAtJ and K fragments.
Fig. 4(b) (and Fig. 4(c)) shows the inverse speed of
ligation v−1
AtJK, respectively)
and the concentration of PEG. We find three distinct
regime in both surface plots: First, at the buffer solu-
tion (cpeg = 0), v−1 proportionally increased with [SJ ]−1
(and [SAtJ]−1). Second, v−1 has local minimum but al-
most constant even by changing [SJ ]−1 (and [SAtJ]−1)
at intermediate concentration of PEG such as 10%,. Fi-
nally, v−1 slows down as [SJ ]−1 increases (and [SAtJ]−1
increases) at highly large concentration of PEG. In par-
ticular, the reaction of long and short DNA in Fig, 4(c)
clearly shows that the production of longer ligated DNA
is suppressed by the addition of excess amount of sub-
strate DNA, as consistent with experimental result.
These distinct regimes can be explained from Eq.(1)
as follows: Given the ligation of short fragments S =
([SK], [SJ ], [SJtK ]) and P = (J,K), and the ligation of
short and long fragments S = ([SK], [SAtJ ], [SAtK]) and
P = (AtJ,K).
5
FIG. 3. Kinetics of DNA ligation in crowding solutions is analyzed by qPCR-bSTAR method. (a to c) The relative amount of
JtK, which was synthesized from ligation of J and K, was plotted in time t at various PEG concentrations. Color represents
the initial concentration of DNA substrate J. (a) 0.0% PEG, (b) 10.0% PEG, (c) 15.0% PEG. (d) 0.0% PEG, (e) 10.0% PEG,
(f) 15.0% PEG. (d to f) The relative amount of AtK, which was synthesized from ligation of AtJ and K, was plotted in time
t at various PEG concentrations. Color represents the initial concentration of DNA fragment AtJ. (g and h) Lineweaver-Burk
plot of ligation reactions. (g) DNA ligation of J and K fragments. Fitting curve is linear function. (h) DNA ligation of short
K fragment and long AtJ fragment. Fitting curve is linear function at 0.0% and 10.0% PEG, but it becomes inverse linear
function at 15.0% PEG. (i) Summary of the effect of molecular crowding on the statistical distribution and kinetics of DNA
ligation. Randomness, mean length hLi, and the kinetic speed v show distinct profiles in various PEG concentrations (from
left to right: dilute, crowded, highly crowded regimes).
• DNA ligation in a buffer solution
• Ligation of two short DNA in a polymer solution
At a dilute solution of small cpeg, the effect of
molecular crowding is absent. DNA do not con-
tribute to trap ligase enzyme from the solution and
Θ ∼ 1 and Ψ/[SK]=const. By considering the DNA
concentration dependence on ligation speed Eq.(1),
v−1 ∼ 1 + Ψ
[SK][SJ ] ∼ [SJ ]−1 (and [SAtJ ]−1).
The molecular crowding (effect 2 in Fig. 4(a)) in-
duces the non-specific binding of ligase onto DNA
at the PEG solution. This non-specific binding de-
creases the freely available enzyme from the so-
lution, and the ratio of trapped enzyme becomes
Θ ∼ [SJ ][SK]. By taking into account this ef-
6
a is the binding constant of partially complementary pair, k′
FIG. 4. Model of partial hybridization pathways in DNA ligation with molecular crowding. (a) Two distinct effects involved
in DNA ligation with molecular crowding. ka is the binding constant of complementary pair, kd is the dissociation constant
of hybridized pair, k′
d is the dissociation constant of partially
hybridized pair, k+1 and k−1 are the binding constant and the dissociation constant between DNA and ligase, respectively. Se
i
is a part of Si including sticky ends. Si is a part of Si which do not include sticky ends. (b) Theoretical ligation speed for
the ligation of two short DNA fragments, K and J. The ligation speed is plotted in the initial DNA concentration [J]0 and the
PEG concentration. (c) Theoretical ligation speed for the ligation of one short DNA fragment K and one long DNA AtJ. The
ligation speed is plotted in the initial concentration of the long DNA [AtJ]0 and the PEG concentration.
fect, the ligation of two short fragments proceeds
at v−1 ∼ 1 + Ψ ∼ const.
• Ligation of long and short DNA in a polymer solu-
tion
Long DNA also traps larger amount of ligase by
non-specific binding Θ ∼ [SAtJ ]2 due to crowding
effect (effect 2 in Fig. 4(a)). Hence the ligation of
one short fragment and one long fragment is v−1 ∼
1 + Ψ[SAtJ ]2
[SK ][SAtJ ] ∼ [SAtJ ].
This mathematical analysis suggests that the attrac-
tive but non-specific binding due to molecular crowding
gives rise to the transition of ligation speeds with non-
trivial length dependence.
Optimal crowding effect for long DNA synthesis
Finally, using the theoretical model based on Eq.(1),
numerical simulations were carried out to verify the con-
centration distribution and the maximization of the long
DNA concentration by ligation of a large number of DNA
strands. The ligation kinetics is described by the inter-
play of the synthesis and the loss of [Si],
d[Si]
dt
= XSn+Sm→Si
vnm − X(i,l)∈P
vil.
(2)
The first term represents the production of Si from all
possible DNA pairing (the ligation of Sn + Sm → Si)
and the second term is the loss as substrates for other
DNA reactions (the ligation of Si + Sl → Si+l). The
joining reaction of 11 kinds of DNA strands was numeri-
cally calculated using the Gillespie method, and the con-
centration distribution at the time when T progressed
for a certain time was plotted against the DNA strand
7
tistical analysis.Heretofore, as a scenario for avoiding the
exponential decay due to random ligation, it has been
discussed that local concentration amplification cancels
the decay by raising the initial concentration[6][7][10]. To
increase the proportion of long DNA strands in a molec-
ular population crowded with various kinds of polymer
solute[25], we found that molecular crowding by a coex-
isting polymer enhances the efficiency of the enzymatic
reaction, in particular DNA ligation.
Furthermore, we have found the stretched exponen-
tial decay of ligated product at highly crowded condition
(15.0% PEG). The deviation from exponential regime im-
plies non-random manner in ligation kinetics. Long-tail
distribution such as power-law[26] is also important to
suppress exponential decay. Next question is to find the
mechanism of enhanced synthesis of long DNA with non-
random manner in simple physicochemical condition. It
is worthy to note that the microscopic attraction be-
tween ligase and DNA is not experimentally detected
in this study though it is one plausible mechanism to
explain the optimal DNA ligation. The interaction of
DNA-binding protein onto DNA has been extensively
discussed[27] and the analysis in detail will be addressed
in future study. Finally, there are crowded conditions
in which high concentrations of proteins are present in
living cells[28]. Whether enzymes react randomly or
non-randomly in these circumstances is little understood.
One of challenges for future studies is to link the dynam-
ics of one enzyme molecule in a living cell with appearing
functions in the whole intracellular environment.
Materials and Methods
1. Synthesis of short DNA fragments
Short DNA fragments used as substrates for lig-
ation reactions were synthesized through polymerase
chain reaction (PCR). DNA from the pUC19 plas-
mid was used as template for the synthesis of 11 dif-
ferent non-overlapping DNA fragments, and a KOD-
plus-Neo polymerase (KOD-401, Toyobo) was used for
DNA amplification. Resulting DNA fragments were
258-bp long and were labeled alphabetically as A, B,
· · · , K. Oligonucleotide primers were designed by us-
ing Primer3web[29][30][31] and were purchased from Eu-
rofins Genomics. The sequences of oligo primers used in
this study are listed in the Supplementary table[23].
Synthesized DNA was treated with a restriction en-
zyme with Bst X1 (R0113L, New England BioLabs) at
37.0 ◦C overnight. Bst X1 recognition sites were added
to the 5′ end of each primer. Purified DNA fragments
have a sticky end at each side, except those synthesized
using the forward primer for the A fragment and the
reverse primer for the K fragment. In addition, the re-
verse primer for the K fragment has a biotin in its 5′ end
in order to capture 5′ biotinylated K fragment for the
pull-down step (Fig.1(a), step4).
FIG. 5. Numerical analysis of stochastic ligation of 11 DNA
species in crowded solutions.
(a) Ligation kinetics of the
number of ligated DNA products. Kinetics of four DNA
species are plotted over time. We define the characteristic
time τ when the first longest DNA AtK was synthesized.
(b) Normalized cumulative distribution function F (L) with
various concentrations of PEG. The color code represents
the concentration of crowding agent. The fitting curves is
F (L) = exp(−αL) with α = 0.893 in 0.0% PEG, α = 0.407
in 3.0% PEG, α = 0.003 in 7.0% PEG, α = 3.662 in 15.0%
PEG. (c) The mean length hLi in the concentration of crowd-
ing agent and the reaction time t (in unit of τ ). The blue
color represents the longer mean length.
length (Fig. 5(a)). As the concentration of PEG was
increased, the slope of F (L) became smaller but, as the
PEG concentration further increased, the slope of F (L)
became steep again. Moreover, this model reproduced
that the concentration of longest DNA (Fig. 5(b)) and
the mean length of ligated DNA hLi (Fig. 5(c)) also
had a peak at intermediate PEG concentrations. On the
one hand, cCrowding effect promotes the joint formation
and increases the rate of ligation per unit time. On the
other hand, because the large fraction of ligase would be
trapped onto DNA at the higher the PEG concentration,
the effective concentration of freely available enzyme be-
comes lower and in turn the ligation reaction slows down.
Such interplay of enhanced binding and reduced enzyme
concentration gives rise to optimality for the production
of long DNA.
DISCUSSION
In this study, we report the enhanced ligation of DNA
in crowded polymer solution by using qPCR-based sta-
2. DNA ligation
The final volume of the ligation reaction solution was
10 µL, where 7.27 ng of 11 short DNA fragments and ther-
mostable Taq DNA ligase (M0208, New England Biolab)
at a final concentration of 20 U were mixed. Ligation
reaction was performed at 25.0 ◦C for 15 h. Short DNA
fragments were designed to be linked in alphabetical or-
der, i.e. A fragment was linked to B fragment, and B
fragment bind to C fragment.
3.
qPCR analysis
The qPCR reaction volume was 10 µL. We mixed 5 µL
of 2×TB Green Premix Ex Taq II (RR820, TaKaRa) with
0.25 µL of extracted DNA-magnetic beads, 0.5 µL of the
forward primer, 0.5 µL of the reverse primer, and 3.75 µL
of MilliQ water. The final concentration for each DNA
primer was 400 nM. Thermal cycling for qPCR analysis
was performed using a real-time PCR system (CFX96,
Bio-Rad) as follows: heating to 95.0 ◦C for 30 s; 35 cycles
of 95.0 ◦C for 10 s and 60.0 ◦C for 30 s. The threshold
cycle number Cq was determined and the initial DNA
concentration was calculated according with this value
as follows: Using pUC19 plasmid DNA as a template
DNA of a known concentration, the Cq value of each
8
primer was plotted into a calibration curve between the
Cq value and DNA concentration
For the measurement of ligation speed vij , we mixed
K fragment (final conc. 48.6 pM) with either J fragment
or AtJ fragment at various concentrations. The ligation
reaction was performed at 25.0 ◦C with a thermostable
Taq DNA ligase (final conc.
6.4 U) and the volume
of reaction mixture was 4 µL. We chose time steps
to observe linear increment of the relative amount of
ligated DNA.
Acknowledgements
This work was supported by Grant-in-Aid for Scientific
Research on Innovative Areas (JP16H00805 Synergy
of Structure and Fluctuation, JP17H05234 Hadean
Bioscience, and JP18H05427 Molecular Engines) and
Grant-in-Aid for Scientific Research (B) JP17KT0025
from MEXT, and Human Frontier Science Program
Research Grant (RGP0037/2015). We thank R. Sakai,
K. Yoshimoto and S. Terada for experimental assistance.
Competing interests
The authors declare no competing interests.
Corresponding address
[email protected]
[1] B. Alberts, A.D. Johnson, J. Lewis, D. Morgan, M. Raff,
K. Roberts, and P. Walter, Molecular Biology of the Cell
(CRC Press, 2017).
[2] R. Phillips, J. Kondev, J. Theriot and H. Garcia, Physical
Biology of the Cell (CRC Press, 2012).
[3] C.A. Hutchison et al., Design and synthesis of a minimal
bacterial genome, Science 351, aad6253 (2016).
[4] C.S. Reddy, A. Arinsteina and E. Zussman, Polymeriza-
tion kinetics under confinement, Polym. Chem. 2, 835-839
(2011).
[5] L.D. Landau and E.M. Lifshitz, Statistical Physics 1
(Butterworth-Heinemann, 1980).
[6] D. Braun and A. Libchaber, Trapping of DNA by ther-
mophoretic depletion and convection, Phys. Rev. Lett. 89,
188103 (2002).
[7] P. Baaske, F.M. Weinert, S. Duhr, K.H. Lemke, M.J. Rus-
sell and D. Braun, Extreme accumulation of nucleotides in
simulated hydrothermal pore systems, Proc. Natl. Acad.
Sci. USA 104, 9346 (2007).
[8] C.B. Mast, S. Schink, U. Gerland and D. Braun, Escala-
tion of polymerization in a thermal gradient, Proc. Natl.
Acad. Sci. USA 110, 8030 (2013).
[9] A. Priye, Y. Yu, Y.A. Hassan and V.M. Ugaz, Synchro-
nized chaotic targeting and acceleration of surface chem-
istry in prebiotic hydrothermal microenvironments, Proc.
Natl. Acad. Sci. USA 114, 1275 (2017).
[10] Y.T. Maeda, A. Buguin and A. Libchaber, Thermal sep-
aration: Interplay between the Soret effect and entropic
force gradient, Phys. Rev. Lett. 107, 038301 (2011).
DNA folding and small RNA stem-loop in thermophoresis,
Proc. Natl. Acad. Sci. USA 109, 17972 (2012).
[12] B.H. Pheiffer, and S.B. Zimmerman, Polymer-stimulated
ligation: Enhanced blunt- or cohesive-end ligation of DNA
or deoxyribooligonucleotides by T4 DNA ligase in polymer
solutions, Nucl. Acid. Res. 11, 7853 (1983).
[13] D. Marenduzzo, K. Finan and P.R. Cook, The depletion
attraction: An underappreciated force driving cellular or-
ganization, J. Cell Biol. 175, 681 (2006).
[14] S. Asakura and F. Oosawa, Interaction between particles
suspended in solutions of macromolecules, J. Polym. Sci.
B 33, 183 (1958).
[15] D. Kilburn, J.H. Roh, L. Guo, R.M. Briber, S.A. Wood-
son and T.C. Jenkins, Molecular crowding stabilizes folded
RNA structure by the excluded volume effect, J. Am.
Chem. Soc. 132, 8690 (2010).
[16] J.S. Kim, V. Backman and I. Szleifer, Crowding-induced
structural alterations of random-loop chromosome model,
Phys. Rev. Lett. 106,168102 (2011).
[17] H. Kang, P.A. Pincus, C. Hyeon and D. Thirumalai,
Effects of macromolecular crowding on the collapse of
biopolymers, Phys. Rev. Lett. 114, 068303 (2015).
[18] B.Akabayov, S.R. Akabayov, S-J. Lee, G. Wagner and
C.C. Richardson, Impact of macromolecular crowding on
DNA replication, Nat. Commun. 4, 1615 (2013).
[19] J. Szymanski and M. Weiss, Elucidating the origin of
anomalous diffusion in crowded fluids, Phys. Rev. Lett.
103, 038102 (2009).
[20] I.M. Sokolov, Models of anomalous diffusion in crowded
[11] Y.T. Maeda, T. Tlusty and A. Libchaber, Effects of long
environments, Soft Matt. 8, 9043 (2012).
9
[21] I. Budin and J.W. Szostak, Expanding roles for diverse
physical phenomena during the origin of life, Ann. Rev.
Biophys. 39, 245 (2010).
[22] D. Kestemont, M. Renders, P. Leonczak, M. Abramov,
G. Schepers, V.B. Pinheiro, J. Rozenski and P. Herdewijn,
XNA ligation using T4 DNA ligase in crowding conditions,
Chem. Commun. 54, 6408 (2018).
[23] See supplemental material for details of experimental
methods, full theoretical details.
[24] L.Homchaudhuri, N. Sarma and R. Swaminathan, Effect
of crowding by dextrans and Ficolls on the rate of alkaline
phosphatase-catalyzed hydrolysis: a size-dependent inves-
tigation, Biopolymers 83, 477 (2006).
[25] A. Blokhuis, D. Lacoste, P. Nghe and L. Peliti, Selection
dynamics in transient compartmentalization, Phys. Rev.
Lett. 120, 158101 (2017).
[26] I. Grossman-Haham, G. Rosenblum, T. Namani and H.
Hofmann, Slow domain reconfiguration causes power-law
kinetics in a two-state enzyme, Proc. Natl. Acad. Sci. USA
115, 513 (2018).
[27] G-W. Li, O.G. Berg and J. Elf, Effects of macromolecular
crowding and DNA looping on gene regulation kinetics,
Nat. Phys. 5, 294 (2009).
[28] A.J. Boersma, I.S. Zuhorn and B. Poolman, A sensor for
quantification of macromolecular crowding in living cells,
Nat. Methods 12, 227 (2015).
[29] A. Untergasser, I. Cutcutache, T. Koressaar, J. Ye, B.C.
Faircloth, M. Remm and S.G. Rozen, Primer3 - new capa-
bilities and interfaces, Nucleic Acids Res. 40, e115 (2012).
[30] T. Koressaar and M. Remm, Enhancements and modifi-
cations of primer design program Primer3, Bioinformatics
23, 1289 (2007).
[31] T. Koressaar, M. Lepamets, L. Kaplinski, K. Raime,
R. Andreson and M. Remm, Primer3-masker: integrating
masking of template sequence with primer design software,
Bioinformatics 34, 1937 (2018).
I. SUPPLEMENTAL MATERIALS AND METHODS
Capillary electrophoresis
10
We also checked the exponential distribution of DNA concentration with its size by capillary electrophoresis. An elec-
trophoresis band as shown in FIG.S1(a) was obtained using an electrophoresis apparatus (Agilent, Bioanalyzer2100).
When the DNA concentration was quantified from the concentration of these bands, the exponential decay was ob-
served (Fig. S1(b)) as same as qPCR-bSTAR analysis. However, in capillary electrophoresis, a band with a DNA
length of 7 or more could not be detected. This result supports the advantage of qPCR-bSTAR analysis in order to
quantitatively measure the distribution function.
(a)
(b)
(c)
FIG. S 1. Capillary electrophoresis assay of ligated DNA. (a) Bands of DNA separated by capillary electrophoresis. Left lane
is marker DNA, middle lane is ligated DNA in a buffer solution, right lane is ligated DNA in 5.0% PEG. (b) Concentration
of DNA bands, which are products of DNA ligation in a buffer solution, with respect to the DNA length. Black dots are
experimental data and blue solid line is fitting curve of exponential function. (c) Concentration of DNA bands, which are
products of DNA ligation in the 5.0% PEG solution, with respect to the DNA length. Black dots are experimental data and
blue solid line is fitting curve of exponential function.
qPCR quantification
As described in main text, short DNA fragments used in end-to-end ligation were made by polymerase chain reaction
(PCR) with KOD-plus-Neo polymerase (KOD-401, Toyobo). We used pUC19 plasmid DNA as template for PCR
amplification and 11 different sequences were selected short DNA fragments. Resulting DNA fragments were 258-bp
long and were labeled alphabetically as A, B, · · · , K. As desbribed in main text, we represent the ordered sequence
of ligated DNA products as XtY, meaning that short DNA fragments from the X fragment to the Y fragment are
connected into one DNA polymer. For instance, a DNA that includes A-B-C-D sequence is referred to as AtD.
While K-K refers to the K fragment alone. We designed qPCR primers that can make hybridization pairing at the
junctions connecting A-B, B-C, · · · , and J-K. Because these qPCR primers also have complementary sequences to
pUC19 plasmd DNA, we can also obtain qPCR growth curve by using pUC19 DNA. This growth curve allows us to
draw a calibration curve between the Cq value with each primer set and the actual DNA concentration (Fig. S2).
This experiment was performed with all qPCR primer sets.
11
Protocol for qPCR-bSTAR
• qPCR-bSTAR Step 1: DNA ligation.
The ligation of 11 short DNA fragments (A, B, · · · , K) was performed in a test tube at a temperature of 25.0 ◦C
for 15 hours (Fig.1(a) in main text, first and second steps). Each short DNA fragment had sticky-ends after
digestion with the restriction enzyme Bst X1. We designed an ordered paring, for instance A fragment binds to
B fragment, and B fragment binds to A fragment at one end and to C fragment at the other end.
• qPCR-bSTAR Step 2: Selection of DNA fragments for analysis.
The ligation of 11 different short DNA fragments in a single tube produces 66 types of ligated DNA products
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h)
(i)
(j)
(k)
FIG. S 2. Calibration curve of threshold cycle Cq and given DNA concentration. (a to k)
(Fig.1(a), third step). Rather than measuring the concentration of each of the 66 types of DNA, we extracted
a subpopulation of the 11 types of DNA molecules that have K sequence in common as AtK, BtK, CtK,
· · · , K, but its length covers all possible size. For this aim, the K fragment was biotinylated at its 3' end.
This modification allowed selecting our target population by using a streptavidin pull-down purification method
(Fig. 1(a), fourth step). We utilized magnetic beads covered with streptavidin, Dynabeads M-270 Streptavidin
(Thermo Fisher Scientific), that bound tightly to the biotin conjugated to the K fragment, for the separation of
our target molecules from the rest. These DNA molecules containing a K-sequence were subjected to subsequent
qPCR analysis.
12
• qPCR-bSTAR Step 3: qPCR analysis.
n=1 c(n) where n is the number of DNA fragments per molecule (Fig.2(a)).
Then, the concentration of each ligation junction and K fragment was measured (Fig.1(b)). The concentration
of DNA detected by qPCR primer set K reported the total concentration of all 11 DNA types; it was defined
In the same way,
the concentration of the ligation junction between the J fragment and the K fragment was defined as C(2) =
n=2 c(n), which reported the total concentration of DNA molecules that have a J-K junction, i.e. AtK,
BtK, · · · , JtK. A more general definition is that the measured concentration of a ligation junction between
the X fragment and the X+1 fragment equals the total concentration from AtK, · · · , to XtK. Therefore, the
n=L c(n). For the
analysis of experimental data, a normalized cumulative distribution function (CDF) F (L) := C(L)/C(1) =
n=1 c(n) was used, because the exponential function is retained if the cumulative distribution
n=L p(n) by using a probability distribution
as C(1) = P11
P11
concentration obtained from qPCR analysis is the cumulative concentration of S(L) = P11
P11
n=L c(n)/P11
function has a large L. CDF F (L) was also expressed as F (L) =P11
function p(n) = c(n)/P11
n=1 c(n).
Oligo DNA primer sequence
Sequences of oligo DNA primer used in the synthesis of 11 short DNA fragments are listed in Supplemental Table
I, and sequences of oligo DNA primer for the qPCR quantification are listed in Supplemental Table II.
We also confirmed that non-specific detection by qPCR primers is negligible (FIG.S3). One primer set is designed
to uniquely detect one DNA sequence. Even if the wrong sequence is read, its error concentration is very small, less
than 10−2 times of the DNA with the correct sequence. This detection error does not affect qPCR-bSTAR analysis.
FIG. S 3. Nonspecific amplification was checked. We measured concentration of each DNA(0.9 pM)(AB,· · · ,K) which has only
one junction by using primer sets shown in TABLE. SII. C0 is concentration of DNA measured by a matched primer set(i.e.
concentration of AB measured by pAB), and C is concentration of DNA measured by other primer sets(i.e. concentration of
AB measured by pBC). NTC means no template control.
fragment direction
sequence(5′ − 3′)
13
GCGCCCAATACGCAAACCGCCTCTC
TABLE S I. The list of oligo primers used for synthesis of short DNA fragments by PCR and their sequences. Sticky ends cut
by restriction enzyme treatment are shown in bold. The target sequence recognized by restriction enzyme are underlined.
[BioON]GTCGTGCCAGCTGCATTAAT
A
B
C
D
E
F
G
H
I
J
K
CCCCAGATCCTTGGCGTCAGGTGGCACTTTTC
CCCCAAGGTACTGGCCCAACTGATCTTCAGCATC
CCCCAGTACCTTGGTGCACGAGTGGGTTACA
CCCCAATGTGCTGGCTGCAGGCATGCAA
CCCCAGCACATTGGGTCGACTCTAGAGGA
CCCCAAGCAACTGGTATGCGGTGTGAAATA
forward
reverse
forward
reverse
forward CCCCAGTTGCTTGGTGGTGCACTCTCAGTACAATC
reverse CCCCAAGGATCTGGTCTAAGAAACCATTATTATCA
forward
reverse
forward
reverse CCCCAAATGCCTGGACTGCATAATTCTCTTACTGT
forward CCCCAGGCATTTGGGCTGCCATAACCATGAGTGAT
reverse CCCCAAATCGCTGGTTAATTGTTGCCGGGAAGCTA
forward CCCCAGCGATTTGGTAGACTGGATGGAGGCGGATA
reverse CCCCAATAGCCTGGGACAGTTACCAATGCTTAATC
forward CCCCAGGCTATTGGAGACCAAGTTTACTCATATAT
reverse CCCCAATACGCTGGTTGATCCGGCAAACAAACCAC
forward CCCCAGCGTATTGGGAGCTACCAACTCTTTTTCCG
reverse CCCCAATGCTCTGGTGCACGAACCCCCCGTTCAGC
forward CCCCAGAGCATTGGCACAGCCCAGCTTGGAGCGAA
reverse CCCCAAACGACTGGCATAGGCTCCGCCCCCCTGAC
forward CCCCAGTCGTTTGGGAAAAACGCCAGCAACGCGGC
reverse
name recognized junction direction
sequence(5′ − 3′)
14
pAB
pBC
pCD
pDE
pEF
pFG
A-B
B-C
C-D
D-E
E-F
F-G
pGH
G-H
pHI
pIJ
pJK
pK
H-I
I-J
J-K
K
TCCGGCTCGTATGTTGTGTG
TTGTAAAACGACGGCCAGTG
CGCATCTGTGCGGTATTTCAC
GCTTGTCTGTAAGCGGATGC
ACGAAAGGGCCTCGTGATAC
TTTCCGTGTCGCCCTTATTC
GGGCGAAAACTCTCAAGGA
TCGCCGCATACACTATTCTCA
GAAGTAAGTTGGCCGCAGTG
CGGAGCTGAATGAAGCCATAC
forward
reverse
forward
reverse
forward
reverse GTCTCATGAGCGGATACATATTTGA
forward
reverse
forward
reverse
forward
reverse
forward CGAAATAGACAGATCGCTGAGATAG
reverse ACTCACGTTAAGGGATTTTGGTCA
forward
reverse
forward
reverse
forward
reverse
forward
reverse
GCGTCAGACCCCGTAGAAA
GCCAGTTACCTTCGGAAAAAGA
ACTTTATCCGCCTCCATCCA
AGTCGTGTCTTACCGGGTTG
TGGCGCTTTCTCATAGCTCA
TTCGCCACCTCTGACTTGA
GCAGGAAAGAACATGTGAGCA
CTTTTGCTGGCCTTTTGCTC
CTTCCTCGCTCACTGACTCG
TABLE S II. The list of oligo DNA primers used for qPCR analysis and their sequences.
fragment direction
sequence(5′ − 3′)
15
GCGCCCAATACGCAAACCGCCTCTC
CCCCAGTACCTTGGTGCACGAGTGGGTTACA
AB
BC
CD
DE
EF
FG
GH
HI
IJ
JK
K
CCCCAGATCCTTGGCGTCAGGTGGCACTTTTC
CCCCAAGCAACTGGTATGCGGTGTGAAATA
CCCCAGCACATTGGGTCGACTCTAGAGGA
forward
reverse
forward
reverse CCCCAAGGATCTGGTCTAAGAAACCATTATTATCA
forward CCCCAGTTGCTTGGTGGTGCACTCTCAGTACAATC
reverse
CCCCAAGGTACTGGCCCAACTGATCTTCAGCATC
forward
reverse CCCCAAATGCCTGGACTGCATAATTCTCTTACTGT
forward
reverse CCCCAAATCGCTGGTTAATTGTTGCCGGGAAGCTA
forward CCCCAGGCATTTGGGCTGCCATAACCATGAGTGAT
reverse CCCCAATAGCCTGGGACAGTTACCAATGCTTAATC
forward CCCCAGCGATTTGGTAGACTGGATGGAGGCGGATA
reverse CCCCAATACGCTGGTTGATCCGGCAAACAAACCAC
forward CCCCAGGCTATTGGAGACCAAGTTTACTCATATAT
reverse CCCCAATGCTCTGGTGCACGAACCCCCCGTTCAGC
forward CCCCAGCGTATTGGGAGCTACCAACTCTTTTTCCG
reverse CCCCAAACGACTGGCATAGGCTCCGCCCCCCTGAC
forward CCCCAGAGCATTGGCACAGCCCAGCTTGGAGCGAA
reverse
forward CCCCAGTCGTTTGGGAAAAACGCCAGCAACGCGGC
reverse
[BioON]GTCGTGCCAGCTGCATTAAT
[BioON]GTCGTGCCAGCTGCATTAAT
TABLE S III. The list of oligo DNA primers used for the error-estimation in qPCR measurement.The alphabet set listed in
the left column is the label sequence (AB, BC, · · · , JK, K) of the synthesized DNA. We used these DNA for the analysis of
uniqueness of the sets of qPCR primer. The experimental result is shown in FIG.S3
II. THEORETICAL MODEL
A. Kinetic model for DNA ligation in crowded solution
16
In this section, we show the derivation of Eq.(1) in maintext where the speed of ligation in crowded condition is
given by the concentration of DNA fragments with sticky ends. First, we
• E : ligase
• Si : DNA i
• [Si] : concentration of DNA i (mol/L)
• Li : length of DNA i (bp)
• lL : length of DNA where ligase search ligation site while ligase attach to DNA (bp)
• SiSj : complex of DNA i and DNA j by complementary base-pairing interactions between adhered cohesive
ends
• ESiSj : complex of ligase and SiSj
• Si,j : DNA which is ligated DNA i and DNA j
• Li,j = Li + Lj
• P = {(i, j) sticky end of DNA i and that of DNA j is complementary.}
• I = {(i, j)sticky end of DNA i and that of DNA j is NOT complementary.}
Consider N species of DNA and a ligation reaction of Si + Sj → Sk. Because the short homologous DNA sequences
make hybridization of double strand DNA structure at the sticky ends, DNA ligase recognizes this hybridized DNA
region and subsequently carries out homologous recombination, thus DNA can be defined as two distinct regions:
• Se
i : homologous lL(bp) sequence at sticky ends of DNA Si which sticks to DNA Sj.
• Si : non-homologous area of DNA Si
Because Si has only 1 Se
i , the concentrations of Se
i and Si holds the following conservation rule
[Se
i ] = [Si].
On the other hand, the length of Si is Li − lL and the concentration of Si can be described as follows
with prefactor M (Si; lL) is given by
[Si] = M (Si; lL)[Si].
M (Si; lL) ≡( Li/lL − 1 (Li/lL ≥ 1)
(Li/lL < 1),
0
(1)
(2)
(3)
By considering these distinct DNA sequences shown above, the DNA ligation reaction is described by the following
chemical reactions((i, j) ∈ P); Hybridization and dissociation of Si and Sj are given by
Si + Sj
ka
⇋
kd
SiSj.
The kinetic constant of association and dissociation of ligase on the hybridized sticky ends is given by
E + (SiSj)e
k+1
⇋
k−1
E(SiSj)e.
(4)
(5)
The binding and dissociation of ligase at homologous sequence of Si that can make hybridization pairing with Sj and
vice versa is given by
17
E + Se
i
k+1
⇋
k−1
ESe
i .
Hybridization of Sj(i) and Si(j) which has E on its homologous sequence are given by
ESe
i(j) + Sj(i)
ka
⇋
kd
E(SiSj)e.
The binding and dissociation of ligase at non-homologous sequence of Si(j) is given by
Finally, the synthesis of ligated product is is given by
E(SiSj)e k+2−−→ E + Sk.
E + Si(j)
k+1
⇋
k−1
ESi(j).
(6)
(7)
(8)
(9)
Chemical reactions shown above are limited to the reaction pathway among two DNA substrate with perfect base
pairing in hybridization sequence, but in order to explain the optimal enhanced DNA ligation in crowded solution,
we additionally assume the imperfect hybridization of two DNA Sl and Sm where their sticky ends do not completely
match as complementary DNA sequence ((l, m) ∈ I);
Such partial hybridization can be induced by the presence of crowding agent given that the attractive interaction in
large polymer arises from the depletion force in crowded polymer solution.
Sl + Sm
ka′
⇋
kd′
SlSm.
(10)
III. THEORETICAL ANALYSIS OF LIGATION SPEED IN CROWDED ENVIRONMENT
Setup for chemical reaction
Next, we derive the speed of ligation of Si + Sj vij ((i, j) ∈ P) by solving Michaelis-Menten equation of the chemical
reactions Eqs. (4) to (10).
The kinetics of the complex synthesis [E(SiSj)e] is
d
dt
[E(SiSj)e] = k+1[E][(SiSj)e] + ka([ESe
i ][Se
j ] + [ESe
j ][Se
i ]) − (k−1 + 2kd + k+2)[E(SiSj)e]
(11)
where (i, j) ∈ P. We have to convert [E(SiSj)e] to C[E][Si][Sj] with constant C. At the steady-state d
Eq. (11) is rewritten by
dt [E(SiSj)e] = 0,
(12)
In order to get the speed of ligation, we have to rewrite [E(SiSj)e] by [E], [Si] and [Sj]. To solve Eq.(12), we need to
obtain explicit forms of [SiSj] and [ESe
i ],at steady state is described by,
i ]. First, according to Eqs. (6)(7)(10), [ESe
(k−1 + 2kd + k+2)[E(SiSj)e] = k+1[E][(SiSj)e] + ka([ESe
j ] + [ESe
j ][Se
i ][Se
i ]).
k+1[E][Se
[ESe
i ] =
i Se
i ] + kdP(i,l)∈P[ESe
k−1 + kaP(i,l)∈P[Se
l ] + kd′P(i,l)∈I[ESe
l ] + ka′P(i,l)∈I[Sl]
i Sl]
.
(13)
For a dilute solution with ∀[Si] ∈ S ≪ 1 higher order terms such as [Si]3 are sufficiently small to assume in Eq.(12)
kaP(i,l)∈P[Se
constant k−1. The approximation of low concentration limit, local balance among [E], [Se
l ] + ka′P(i,l)∈I[Sl] ≪ k−1. One can find that the denominator in the right side of Eq. (13) becomes
i ], and [ESe
i ] is given by
[ESe
i ] =
[E][Se
i ].
k+1
k−1
The same procedure also leads the local balance among [(SiSj)e], [Se
i ], and [Se
i ][Se
j ].
kd[(SiSj)e] = ka[Se
j ] from Eq(4) as
By using Eqs. (14) and (15), Eq. (12) is simplified as
This expression is known as Lineweaver-Burk plot for enzymatic reaction in particular for DNA ligation.
[E(SiSj)e] = [E]
k+1ka
k−1 + 2kd + k+2(cid:16) 1
kd
+
1
k−1(cid:17)[Se
i ][Se
j ].
(14)
(15)
(16)
The reaction pathway of DNA ligation in crowded solution
Next, we give the explicit form of active ligase concentration [E] as a function of total amount of ligase and substrate
DNA concentrations in order to solve the speed of DNA ligation in crowded environment. Because the total amount
of the enzyme is constant [E0], the fraction of the complexes made of ligase and DNA is given by
18
[E]0 = [E] + X(l,m)∈P
([E(SlSm)e] + [ESlSm]) + X(l,m)∈I
[ESlSm] +Xk
([ESe
k] + [ESk])
(17)
where [E] is the freely available ligase in solution, P(l,m)∈P is the summation over all possible complementary
pairing of DNA, and P(l,m)∈I is the summation over all partial complementary pairing of DNA
Given that the association/dissociation kinetics of ligase and DNA strands is faster than the chemical reaction,
each terms for the ligase-DNA complexes in Eq.(17) are rewritten as
([ESe
n] + [ESn]) =
Xn
=
k+1
k−1
k+1
k−1
[E]Xn
[E]Xn
([Se
n] + [Sn])
(M (Sn; lL) + 1)[Sn]
(18)
from Eq. (8)(14).
Using Eq.(8)(9), the kinetic production of complex [ESlSm]((l, m) ∈ I) in Eq.(18), which is complex of ligase and
partial complementarily hybridized DNA pairing, is described by,
d
dt
[ESlSm] = ka′ ([ESl][Sm] + [Sl][ESm]) + k+1[E](M (SlSm; lL) + 1)[SlSm] − (k−1 + 2kd′)[ESlSm]
= k+1ka′ [E]{k−1
−1(M (Sl; lL) + M (Sm; lL) + 2) + k−1
d′ (M (SlSm; lL) + 1)}[Sl][Sm]
−(k−1 + 2kd′)[ESlSm].
(19)
because we dont have to consider homologous sequences of Sm and Sl ((l, m) ∈ I), length of Sm = M (Sm; lL) + 1.
Assuming steady-state of d
dt [ESlSm] = 0, Eq. (19) is simplified as,
[ESlSm] =[E]
k+1ka′
k−1 + 2kd′
X(l,m)∈I
{k−1
−1(M (Sl; lL) + M (Sm; lL) + 2) + k−1
d′ (M (SlSm; lL) + 1)}[Sl][Sm]
× X(l,m)∈I
(20)
On the other hand, the complex [ESiSj]((i, j) ∈ P) is a complex of ligase [E] and an non-homologous region of SiSj.
From Eq.(8)(4), its kinetics is represented by
d
dt
[ESiSj] = k+1[E]M (SiSj; lL)[SiSj] + ka([ESi][Sj] + [Si][ESj]) − (k−1 + 2kd)[ESiSj]
= [E]k+1kan 1
kd
−(k−1 + 2kd)[ESiSj].
M (SiSj; lL) +
1
k−1
(M (Si; lL) + M (Si; lL))o[Si][Sj]
(21)
By taking similar approximation if the system has reached steady state, that is d
simplified as
dt [ESiSj] = 0, Eq. (21) is also
[ESiSj] = [E]
M (SiSj; lL) +
k+1ka
k−1 + 2kdn 1
kd
By substituting [ESiSj]((i, j) ∈ P)(Eq.(22)), [ESlSm]((l, m) ∈ I)(Eq.(20)) and [ESe
n] + [ESn](Eq.(18)) into Eq. (17),
1
k−1(cid:16)M (Si; lL) + M (Si; lL)(cid:17)o[Si][Sj].
(22)
−1(M (Sl; lL) + M (Sm; lL) + 2) + k−1
d′ (M (SlSm; lL) + 1)o[Sl][Sm]
−1(cid:16)M (Si; lL) + M (Si; lL)(cid:17)+k−1
d M (SiSj; lL)o[Si][Sj]
the freely available ligase concentration is
[E] = ([E]0 − X(l,m)∈P
[E(SlSm)e])
k+1ka′
+
k+1ka
k−1 + 2kd′ X(l,m)∈Ink−1
×n1 +
k−1 + 2kd X(i,j)∈Pnk−1
k−1 Xn (cid:16)(M (Sn; lL) + 1)[Sn](cid:17)o−1
k+1
+
[E(SlSm)e])Θ(S)−1
= ([E]0 − X(l,m)∈P
k+1ka′
k+1ka
k−1 + 2kd′ X(l,m)∈Ink−1
k−1 + 2kd X(i,j)∈Pnk−1
k−1 Xn (cid:16)(M (Sn; lL) + 1)[Sn](cid:17).
k+1
+
+
where
Θ(S) ≡ 1 +
−1(M (Sl; lL) + M (Sm; lL) + 2) + k−1
d′ (M (SlSm; lL) + 1)o[Sl][Sm]
−1(cid:16)M (Si; lL) + M (Si; lL)(cid:17)+k−1
d M (SiSj; lL)o[Si][Sj]
By inserting Eq.(23) into Eq.(16), we obtain
[E(SiSj)e] = Ψ−1[Se
i ][Se
[E(SlSm)e])Θ(S)−1
j ]([E]0 − X(l,m)∈P
where
Ψ−1 ≡
k+1ka
k−1 + 2kd + k+2
(
1
kd
+
1
k−1
).
The summation of eq.(25) over all possible complementary pairing of DNA is given by,
X(i,j)∈P
[E(SiSj)e] = Ψ−1Θ(S)−1([E]0 − X(l,m)∈P
[E(SlSm)e]) X(i,j)∈P
[Se
i ][Se
j ].
We solve this equation and obtain the following expression
Substituting Eq. (1)(28) into Eq. (25) gives
[E(SiSj)e] =
[E(SlSm)e] =
X(l,m)∈P
[E]0P(i,j)∈P[Se
ΨΘ(S) +P(i,j)∈P[Se
i ][Se
j ]
i ][Se
j ]
[E]0[Si][Sj]
ΨΘ(S) +P(l,m)∈P[Sl][Sm]
.
Finally, the speed of producing Sk from Si and Sj is given by
vij (S) ≡
d
dt
[Sk] = k+2[E(SiSj)e] =
where vij (S) is the speed of ligation of Si + Sj → Sk.
k+2[E]0[Si][Sj]
ΨΘ(S) +P(l,m)∈P[Sl][Sm]
19
(23)
(24)
(25)
(26)
(27)
(28)
(29)
(30)
20
IV. DEPENDENCE OF KINETIC PARAMETERS ON THE CONCENTRATION OF CROWDING
AGENT
In this section, we present theoretical description of the dependence on the concentration of PEG, cpeg.
First, we consider the crowding induced change of the affinity between DNA fragments. The thermodynamic
stability of DNA duplex is important to hybridize sticky ends of DNA fragments. It has been known that crowding
agents such as PEG decreases a melting temperature, Tm, of short DNA fragment [1]. The fraction of duplex DNA
is decided by the thermodynamic stability that decreases at Tm. Therefore the fraction of duplex DNA, r(T, cpeg), is
given by Crowding effects enhance the binding affinity of DNA fragments by entropic force. According to previous
study by Minton, et al[2], the activity of enzyme is given by the following equation as a function of the polymer
concentration,
r(T, cpeg) =
1
1 + exp((T − (Tm − cpeg/3))/10)
,
(31)
where T is temperature in bulk solution and Tm is melting temperature of complementary 4bp of sticky end of DNA[4]
and the dissociation constants KdandK ′
d in the hybridization of sticky ends are proportional to [Si][Sj]/[SiSj] ∼
1/r(T, cpeg). By using this relation, those kinetic constants are reduced to
Kd = 104 ×
1 + exp((T − (Tm − cpeg/3))/10)
1 + exp((T − Tm)/10)
exp(−(cpeg)2)
Kd′ = 106 ×
1 + exp((T − (T ′
m − cpeg/3))/10)
1 + exp((T − T ′
m)/10)
exp(−(cpeg)2)
(32)
(33)
m melting temperature of non-complementary 4bp of sticky end.
In this study, T =25 ◦C, Tm=12 ◦C and
with T ′
T ′
m=4 ◦C.
Both DNA hybridization and DNA-ligase complex formation are given by number density of chemical species, and
the dissociation is regulated by interaction forces. Therefore, the association constant ka and ka′ , and the dissociation
constants kd and kd′ are given by
ka = 100
kd = Kdka
ka′ = 100
kd′ = Kd′ka′ .
(34)
(35)
(36)
(37)
Second, the affinity constant between DNA and ligase is considered Macromolecular crowding suppresses the dis-
sociation of enzyme from bound DNA. The binding constant k+1 and the dissociation constant k−1 are given by
k+1 = 100
k−1 = k+1102 exp(−(cpeg/10)2).
(38)
(39)
Finally, we give explicit form of the production rate of ligated DNA k+2 .It has been known that Taq DNA ligase
catalyze dehydration reaction with coenzyme. In contrast, crowding agent such as PEG attracts water molecules due
to its hydrophilic nature, and the ligation reaction slows down because the fraction of freely available water could
be reduced in higher PEG concentration. We assume that the speed of ligation decreases exponentially with the
concentration of PEG[3], the kinetic constant k+2 for ligation is defined by
k+2 = 103−
cpeg
2
.
(40)
[1] S. Nakano, H. Karimata, T. Ohmichi, J. Kawakami and N. Sugimoto, The effect of molecular crowding with nucleotide
length and cosolute structure on DNA duplex stability, J. Am. Chem. Soc. 126, 14330 (2004).
[2] A.P. Minton, The effect of volume occupancy upon the thermodynamic activity of proteins: some biochemical consequences,
Mol. Cell. Biochem. 55, 119 (1983).
[3] L. Homchaudhuri, N. Sarma and R. Swaminathan, Effect of crowding by dextrans and Ficolls on the rate of alkaline
phosphatase-catalyzed hydrolysis: a size-dependent investigation, Biopolymers 83, 477 (2006).
[4] J. Marks, C. W. Schmid and V. M. Sarich, DNA hybridization as a guide to phylogeny: Relations of the Hominoidea, J.
Hum. Evol. 17, 769 (1988).
|
1108.5074 | 1 | 1108 | 2011-08-25T13:02:43 | Investigating Biological Matter with Theoretical Nuclear Physics Methods | [
"physics.bio-ph",
"hep-lat",
"hep-ph",
"nucl-th",
"q-bio.BM"
] | The internal dynamics of strongly interacting systems and that of biomolecules such as proteins display several important analogies, despite the huge difference in their characteristic energy and length scales. For example, in all such systems, collective excitations, cooperative transitions and phase transitions emerge as the result of the interplay of strong correlations with quantum or thermal fluctuations. In view of such an observation, some theoretical methods initially developed in the context of theoretical nuclear physics have been adapted to investigate the dynamics of biomolecules. In this talk, we review some of our recent studies performed along this direction. In particular, we discuss how the path integral formulation of the molecular dynamics allows to overcome some of the long-standing problems and limitations which emerge when simulating the protein folding dynamics at the atomistic level of detail. | physics.bio-ph | physics |
Investigating Biological Matter with Theoretical
Nuclear Physics Methods
Pietro Faccioli
Dipartimento di Fisica, Universit`a degli Studi di Trento, Via Sommarive 14, Povo (Trento),
I-38123 Italy.
INFN, Gruppo Collegato di Trento, Via Sommarive 14, Povo (Trento), I-38123 Italy.
E-mail: [email protected]
Abstract. The internal dynamics of strongly interacting systems and that of bio-polymers
such as proteins display several
important analogies, despite the huge difference in their
characteristic energy and length scales. For example, in all such systems, collective excitations,
cooperative transitions and phase transitions emerge as the result of the interplay of strong
correlations with quantum or thermal fluctuations.
In view of such an observation, some
theoretical methods initially developed in the context of theoretical nuclear physics have been
adapted to investigate the dynamics of biomolecules.
In this talk, we review some of our
recent studies performed along this direction. In particular, we discuss how the path integral
formulation of the molecular dynamics allows to overcome some of the long-standing problems
and limitations which emerge when simulating the protein folding dynamics at the atomistic
level of detail.
1. Introduction
Proteins are macromolecules involved in virtually all the biochemical processes which take
place inside the cell, such as e.g. catalysis, charge transport, signal transduction. Clearly,
understanding the physical principles which drive the internal dynamics of individual proteins
and shape the protein-protein interaction would have countless implications in molecular biology,
drug design and nano-biotechnology.
In particular, a central open challenge at the interface
of physics, biochemistry and molecular biology concerns the prediction of the protein folding
pathways i.e. the sequence of conformational transitions which these molecules perform in order
to reach their stable and biologically active configuration (see e.g. Fig. 1) .
From a theoretical perspective, the problem of determining the time evolution of proteins has
been mostly investigated by means of Molecular Dynamics (MD) simulations, i.e. through the
direct integration of the classical equations of motion of all the atoms in the system. From the
conceptual point of view, such an approach should provide a reasonable approximation, since
quantum corrections are expected to be quite small. On the other hand, from the practical
point of view, the applicability of classical MD simulations is strongly limited by their intrinsic
computational complexity. Indeed, even using computational resources comparable with those
employed in the largest contemporary lattice QCD simulations, it is possible to follow the
dynamics over time intervals which do not exceed a few hundreds ns [1]. Unfortunately, the
dynamics which drives the protein phenomenology occurs at time scales which are several orders
of magnitude larger, typically ranging from few ms to several seconds.
Figure 1. Schematic representation of a protein folding pathway. This trajectory corresponds
to a biological instanton and was obtained using the Dominant Reaction Pathways approach
(see discussion in section 4.
The decoupling of the microscopic time scales from the time scales at which conformational
reactions occur is a common feature in many bio-molecular systems.
It is due to the fact
that most conformational transitions are cooperative and involve the overcoming of free-energy
barriers. As a result, in a MD simulation, most computational time is wasted to describe thermal
oscillations inside the initial meta-stable state.
Given the huge gap between the time scales at which the relevant physics takes place and
the time intervals which can be covered by MD simulations, it is natural to expect that a real
breakthrough in our comprehension of biological matter at the fundamental level is more likely
to come from new ideas and new theoretical approaches, rather than from the next several
generations of supercomputers.
In the next sections, we discuss how functional integral approaches originally developed
in the context of nuclear theory can provide an extremely powerful framework to investigate
the dynamics of proteins as well as other complex molecular systems. Before introducing the
mathematical details of this theory, it is instructive to discuss some basic aspects of protein
structure and phenomenology and to highlight several analogies with the dynamics of strongly
interacting systems.
2. QCD and Protein Dynamics: So Far Yet So Close
From the biochemical point of view, proteins are poly-peptide chains, i.e. weakly branched
polymers formed by 20 different types of monomers called amino-acids. Typical globular proteins
consist of about 100 amino-acids, but some chains can contain as many as ∼ 30, 000 of such
monomers.
From the physical point of view [2], proteins are very different from essentially all other
known polymers: indeed, while standard hetero-polymers are prototypes of disordered mesoscale
systems, proteins are characterized by their unique "ability" to spontaneously and reversibly fold
into a well-defined stable configuration (denominated the native state), in which they perform
their biological functions (see e.g. Fig. 2). Such a remarkable feature is the result of the
evolutionary pressure, which has selected only a very specific subset in the space of all possible
native stateFigure 2. Ball-and-stick representation of the native state of a globular protein. The tube
represents the back-bone of the chain.
amino-acid sequences.
The native state of a protein is characterized by an extremely low entropy, since it is defined
by the small thermal oscillations around a single conformation. In this sense, proteins may be
regarded as non-symmetric mesoscale crystals. If the temperature of the water surrounding the
molecule is increased, or if a denaturant agent (typically urea) is added to the solvent, then
proteins can reach a much more entropic phase called the molten globule.
In this state, the
chain is still collapsed, but is free to explore the different compact configurations, in analogy
with a liquid drop. Finally, at even higher temperatures or denaturant concentrations, proteins
swell and start to behave as flexible random coils (see right panel of Fig. 3).
Interestingly, the phase diagram of proteins in solution displays several analogies with that
of strongly interacting matter (see left panel of Fig. 3).
Indeed, the low-temperature and
low-density phase of QCD is characterized by a very low entropy, since only colorless hadrons
are present in the spectrum. On the other hand, above the deconfinement phase transition,
the entropy is much larger, since also quantum states in color multiplets can contribute to the
partition function.
In addition, flavor symmetry implies that the hadronic spectrum can be approximatively
organized into SU(3) multiplets. Hence, since quantum states belonging to the same multiplet
can be transformed into one another by a flavor rotation, the number of truly distinct hadronic
wave-functions is actually much smaller than the number of hadrons recorded in the particle
data book. Similarly, it has been experimentally observed that nearly identical native structures
are adopted by poly-peptide chains characterized by very different amino-acid sequences. This
is to say that the number of native structures which can be realized is much smaller than the
number of sequences in the data bank of known proteins.
These analogies are not accidental. They reflect the fact that the dynamics of both hadrons
and proteins is characterized by the interplay of large correlations and quantum or thermal
fluctuations.
It is natural to address the question whether some of the powerful theoretical
techniques developed to describe the non-perturbative internal dynamics of hadrons may be
exported to describe the non-perturbative internal dynamics of proteins. In the next section,
we shall see that this intuition can be made rigorous, i.e. that it is possible to formulate the
dynamics of biomolecular systems in terms of an effective quantum many-body problem.
Figure 3. Comparison between the phase diagram of QCD and of proteins in solution.
3. Path Integral Formulation of Molecular Dynamics
The dynamics of biomolecular systems in solution is often described by means of the Langevin
equation, which reads
m x = −γ x − ∇U (x) + ξ(t),
(1)
In this equation, x = (x1, x2, ..., xN ) is a macro-vector specifying the coordinates of all the
atomic nuclei in the molecule, U (x) is the potential energy and ξ(t) is a stochastic force, which
simulates the effects of the random Brownian collisions with the molecules in the solvent. The
contemporary models for the potential energy U (x) -- which is often improperly called the force
field -- are defined phenomenologically in order to reproduce the results of quantum chemistry
calculations of the equilibrium configurations of small molecules [3]. For sake of simplicity and
without loss of generality, in the following we present the formalism in the case of one particle
in one spacial dimension. The generalization to the multi-dimensional case is straightforward.
The acceleration term on the left hand side of Eq. (1) introduces inertial effects which are
damped on a time-scale t ∼ m/γ ≡ τD. For most biomolecular systems, τD is of order of the
fraction of a ps, hence much smaller than the relevant microscopic time-scale associated to the
dynamics of the torsional angles, which takes place at the ns time-scale.
For t (cid:29) τD the Langevin Eq. (1) reduces to
= − D
kBT
∂x
∂t
∇U (x) + η(t),
(2)
where D = kBT
that satisfy the so-called fluctuation-dissipation relationship:
mγ is the so-called diffusion coefficient and η(t) = 1
γ ξ(t) is a rescaled white noise
It can be shown -- see e.g. [4] -- that the probability distribution sampled by the stochastic
differential Eq. (2) satisfies then the Fokker-Planck (FP) equation:
P (x, t) = D∇
∂
∂ t
+ D ∇2 P (x, t).
(4)
(cid:104)η(t)η(t(cid:48))(cid:105) = 2Dδ(t − t(cid:48)).
(cid:19)
(cid:18) 1
∇ U (x) P (x, t)
kBT
(3)
A universal property of all the solutions of the Eq. (4) is that, in the long time limit, they
converge to the Boltzmann weight, regardless of their initial conditions.
temperature
temperature
chemical
poten/al
denaturant
concentra/on
(urea)
hadronic
phase
deconfined
phase
CFL
phase
folded
phase
molten
globule
phase
coil
phase
PHASE
DIAGRAM
OF
QCD
PHASE
DIAGRAM
OF
PROTEINS
By performing the substitution
P (x, t) ≡ e
− 1
2kB T U (x) ψ(x, t),
the FP Eq. (4) can be formally re-written as a Schrodinger Equation in imaginary time:
− ∂
∂t
ψ(x, t) = Heff ψ(x, t),
where the effective "Quantum Hamiltonian" operator reads
Heff = − D∇2 + Veff (x),
and Vef f (x) is an effective potential and is defined as
(cid:104)
(cid:105)
(∇U (x))2 − 2kBT ∇2U (x)
.
Veff (x) =
D
4(kBT )2
(5)
(6)
(7)
(8)
Hence, we have shown that the problem of studying the real time stochastic dynamics of a protein
in solution can be mapped into the problem of determining the imaginary time evolution of an
effective quantum many-body system.
Based on the analogy with quantum mechanics, it is immediate to obtain a path-integral
representation of the solution of (4), subject to the boundary conditions x(0) = xi and x(t) = xf :
P (xf , txi ) = e
−(cid:82) t
2kB T (U (xf )−U (xi))(cid:104)xie− Heff txf(cid:105)
− 1
− 1
Dx(τ ) e
2kB T (U (xf )−U (xi)) (cid:90) xf
0
= e
(cid:16) x2(τ )
(cid:17)
d τ
4D +Veff [x(τ )]
.
(9)
xi
This equation expresses the fact that the Green's function of the Fokker-Planck equation is
formally equivalent to a quantum-mechanical propagator in imaginary time. On the other hand,
it is important to stress that the variable t which enters in these equations corresponds to the
physical time.
4. The Theory of Biological Instantons
The path integral formulation of the stochastic dynamics of protein is particularly convenient if
one is interested in investigating the structure of the protein folding pathways. Indeed, such a
formalism allows to focus directly on the non-equilibrium dynamical trajectories which actually
reach the native state xf in a given time interval t. This way, one avoids wasting time to simulate
trajectories which remain confined in the initial denatured state. A second important advantage
of Eq. (9) is that it assigns a probability density to each folding pathway x(τ ):
Prob.[x(τ )] ∝ exp[−Sef f [x]].
(10)
Clearly, the folding pathways which are most likely to be realized by a given protein are the
solutions of the "classical" equations of motion generated by the effective action
(cid:90) t
0
(cid:32)
x2
4D
(cid:33)
Sef f [x] =
dτ
+ Vef f [x]
.
(11)
These trajectories typically overcome barriers of the effective potential Vef f (x). Hence, they
correspond to instantons in the effective quantum theory. The so-called Dominant Reaction
Pathways (DRP) approach developed by our group [5, 6, 7, 8, 9] is based on the saddle-point
approximation of the stochastic path integral (9) and focuses on such "biological instantons".
Unlike in QCD, such an approach is not affected by infrared divergences, hence it is completely
rigorous. The expansion parameter controlling the accuracy of this approximation is the thermal
energy kBT , which enters in the definition of the diffusion constant D and of the effective
potential Vef f (x).
In principle,
the instantons can be obtained by relaxing numerically a discretized
representation of the effective action functional Sef f [x], in complete analogy with the cooling
procedure developed in lattice field theory. In practice, however, the simultaneous presence of
fast and slow internal time scales (which range from fs to ns) makes such a task very challenging.
Indeed, the discretized elementary time interval ∆t has to be chosen much smaller than the
smallest time scale. Hence, computing a single dominant pathway would require to find a
minimum of a function with a gigantic number of degrees of freedom: 3N × Nt, where N is the
number of atoms in the protein -- typically from several hundreds to many thousands -- and Nt
is the number of time discretization steps -- typically 106 -- .
Fortunately, a major numerical simplification of this problem can be achieved by exploiting
it
the fact that the dynamical system defined by the effective action (11) is simplectic, i.e.
conserves the "effective energy" 1
Eef f =
1
4D
x2(t) − Vef f [x(t)].
(12)
Hence, rather than minimizing directly the effective action Sef f [x], it is possible to obtain the
biological instantons using the Hamilton-Jacobi (HJ) formulation of classical mechanics. In other
words, the trajectories obeying the classical equations of motion and subject to the boundary
conditions x(t) = xf , and x(0) = xi are those which minimize the effective HJ functional
dl
Eef f (t) + Vef f [x(l)],
(13)
SHJ [x(l)] =
1√
D
(cid:113)
(cid:90) xf
xi
√
dx2 is the measure of the distance covered by the system in configuration space,
where dl =
during the transition. Hence, the dominant reaction pathways can be viewed as the geodesic in
a space with a curved metric.
The major advantage of the HJ formulation is that it allows to remove the time as the
independent variable and replacing it with the curvilinear abscissa l, which has the dimension of
a length scale. Since there is no gap in the length scales of molecular systems, the discretization of
the HJ is expected to converge extremely much faster than the time discretization of the effective
action Sef f [x]. Indeed, typically 100 slices are sufficient to achieve an accurate representation
of the path.
The effective energy Eef f is an external parameter which determines the time at which
each configuration of a dominant reaction pathway ¯x(τ ) is visited, according to the usual HJ
relationship
(cid:90) x
xi
(cid:113)
t(x) =
dl
1
4D(Eef f + Vef f [¯x(l)])
.
(14)
Typically, one is interested in studying transitions which terminate close the local minima of
the potential energy U (x). The residence time in such end-point configurations must be much
longer than that in the configurations visited during the transition. From Eq. (14) it follows
1 Note that Eef f has the dimension of a rate, therefore this quantity does not have the physical interpretation
of a mechanical energy
that these conditions are verified if Eef f ∼ −Vef f (xo), where xo is a configuration in the vicinity
of a local minimum of U (x).
In practice, computing the dominant reaction pathway connecting two given configurations
xi to xf amounts to minimizing a discretized version of the effective HJ functional:
Sd
HJ [x(l)] =
[Eef f + Vef f (x(n))] ∆ln,n+1,
(15)
where Ns is the number of path discretization slices. Once such a path has been determined,
one can reconstruct the time at which each of the configurations is visited during the transition,
using Eq. (14). In particular, the time interval between the n-th and the (n + 1)-th slice is
(cid:114) 1
Ns−1(cid:88)
D
n=1
∆ln+1,n
4D(Eef f + Vef f [¯x(n)])
,
(16)
(cid:113)
where ∆ln+1,n =(cid:112)(¯x(n + 1) − ¯x(n))2.
∆tn+1,n =
4.1. Thermal Fluctuations Around the Dominant Folding Pathways
The biological instantons encode information about the reactive dynamics, hence can be used as
starting point to compute the folding rate, whose inverse gives the typical time a protein takes
to reach the native state, starting from a denatured state. This is a fundamental observable in
protein kinetics, which is accessible from fluorescence experiments.
It turns out that, in order to obtain an equation for the rate, it is necessary to go beyond
the lowest-order saddle-point approximation, and estimate the contribution of small thermal
fluctuations around the dominant paths [8].
This corresponds to evaluating the instanton weight at one-loop level, by functionally
expanding the effective action to quadratic order, around the instanton solution x(τ ):
Sef f [x] = Sef f [x] +
(cid:39) Sef f [x] +
1
2
1
2
(cid:90) t
(cid:90) t
0
dτ(cid:48)(cid:90) t
(cid:90) t
0
dτ
0
0
(xi(τ(cid:48)) − xi(τ(cid:48))) (xk(τ ) − xk(τ )) + . . .
ik [x] (xk(τ ) − xk(τ )),
(17)
dτ(cid:48)(cid:48)
δ2Sef f [x]
δxi(τ(cid:48))δxk(τ(cid:48)(cid:48))
dτ(cid:48)(xi(τ(cid:48)) − xi(τ(cid:48))) F τ,τ(cid:48)
(cid:34)
(cid:35)
where we have introduced the fluctuation operator F [x], defined as
ik [x] ≡ δ2Sef f [x]
F τ,τ(cid:48)
δxi(τ(cid:48))δxk(τ(cid:48)(cid:48))
=
− 1
2D
δik
d2
dτ 2 + ∂i∂k Vef f [x(t)]
δ(τ − τ(cid:48)).
The formal one-loop expression for the conditional probability is therefore
(cid:113)
P (xf , txi) (cid:39) N e− β
2 (U (xf )−U (xi))
det F [x]
e−Sef f [x].
(18)
(19)
Such an expression is clearly divergent and needs to be regularized. To this end, one can
multiply and divide by the conditional probability density of a reference system Preg.(xf , txi).
For example, one can use as regulator the conditional probability associated to the diffusion in
an external harmonic potential, for which the propagator Preg.(x0, tx0) is analytically known.
The result is [8]:
(cid:17) ,
PDRP (xf , txi) = Preg.(x0, tx0) e− β
2 (U (xf )−U (xi)) eSreg.
ef f [xreg.]−Sef f [x]
(cid:118)(cid:117)(cid:117)(cid:116)
1
(cid:16) F −1
det
reg.[xreg.] F [x]
(20)
(cid:16) Freg.[xreg.] F −1[x]
(cid:17)
where Freg is the fluctuation operator for the reference system used in the regularization.
In practice, the determinant det
has to be evaluated numerically from
the discretized representation of the fluctuation operators. To obtain such a representation, one
needs to express the time derivative using discretized time intervals. It is most convenient to
use the intervals ∆ti,i+1 evaluated from the dominant trajectory, according to Eq. (16). The
fluctuation operator reads
−1/D
(cid:32)
(cid:33)
(cid:35)
(cid:34)
1
1
δi,j
δk,m+1
∆tm+1,m
− δk,m
∆tm+1,m
∆tm,m−1
+
+
δk,m−1
∆tm,m−1
F [¯x]i,j
k,m =
∆tm+1,m + ∆tm,m−1
∂2Vef f (¯x(k))
δk,m,
∂xi∂xj
+
(21)
where the indexes k, m = 1, . . . , Ns run over the path frames, while the indexes i, j = 1, . . . , d
label the degrees of freedom of the system.
4.2. Improved Effective Actions
The main numerical advantage of the DRP formalism arises from the possibility of removing the
time as a dynamical variable, and replace it with the curvilinear abscissa l. Since there is no gap
in the characteristic length scales of molecular systems, the convergence of the discretization
of l is usually very fast compare to that of the discretization of t. As a result, using such
a formulation, it is possible to gain information about the reaction mechanism at a very low
computational cost.
On the other hand, in order to obtain information about the dynamics, one needs to compute
the times at which each configuration is visited along the dominant path, using Eq.s (14) and
(16). The calculation of the time intervals ∆ti,i+1 is also needed in the discretized representation
of the fluctuation operator (21), which enters in the calculation of the order kBT corrections
arising from non-equilibrium stochastic fluctuations around the dominant path.
Clearly, the trick to remove time by switching to the HJ formulation does not help in the
evaluation of explicitly time-dependent observables. As a result, in order to achieve an accurate
description of the dynamics, or in order properly take into account of the effects of fluctuations,
one would need to use a large number of path frames, with a consequent significant increase of
the computational cost of a DRP simulation.
Fortunately, the convergence of the calculation of the time intervals in the DRP approach can
be greatly improved by adopting the effective stochastic theory (EST) developed in [10]. The
main idea of the EST is to exploit the gap in the internal time scales in order to analytically
perform the integral over the fast Fourier components of the paths x(τ ) which contribute to the
path integral (9), using Wilson's approach. Through such a procedure, the effects of the fast
dynamics is rigorously and systematically integrated out by renormalizing the effective potential:
Vef f (x) → V EST
ef f (x) = Vef f (x) + V R
ef f (x).
where, to leading order in the effective theory one has
D∆tc (1 − b)
V R
ef f (x) =
2π2b
∇2Vef f (x) + . . .
(22)
(23)
In such an Eq., ∆tc is a cut-off time scale which must be chosen much smaller than the fastest
internal dynamical time scale and b is a parameter which defines the interval of Fourier modes
which are being analytically integrated out. Typically, for molecular systems ∆tc ∼ 10−3ps and
b ∼ 10−2 [11]. The dots in Eq. (23) denote higher order correction in an expansion in the ratio
of slow and fast time scales (so-called slow-mode perturbation theory) and can be found in the
original publication [10].
The EST generates by construction the same long-time dynamics of the original "bare"
theory, but has a lower time resolution. In the context of MD simulations, this implies that the
EST can be integrated using much larger discretization time steps [11]. In the context of the
DRP simulations, the utility of the EST resides in the fact that much fewer path discretization
time steps are required in order to achieve a convergent calculation of the time interval from the
dominant path, through Eq. (14).
4.3. Generalization to Ab-Initio Quantum Chemistry Calculations
So far we have discussed the DRP formalism in the specific context of protein folding. On the
other hand, this theory applies in general to any physical system described by the over-damped
Langevin Eq. (2). In particular, it can be applied to investigate rare chemical reactions taking
place in solution [12]. Typically, such reactions involve large modifications of the electronic
structure, e.g. associated to the breaking or forming of covalent bonds. Hence, in investigating
such processes, a classical description based on some phenomenological expression for the
molecular potential energy U (x) entering in the Langevin equation is no longer appropriate,
and one must adopt a sub-atomic approach, in which the quantum electronic degrees of freedom
are explicitly taken into account.
In the Born-Oppenheimer approximation [3], this is done in two steps. First, the Schrdinger
equation for the electronic degrees of freedom is (approximatively) solved holding fixed all the
nuclear coordinates. The corresponding ground-state energy is then interpreted as the potential
molecular energy U (x) and is used to derive the effective potential Vef f (x), which enters the
DRP formulas. This approach allows to rigorously take into account of the quantum effects on
the electron dynamics, to all order in ¯h and to leading order in the ratio m/M between the mass
of the electron m and the typical mass of the atomic nuclei M .
4.4. Quantum Corrections to the Diffusive Dynamics of the Atomic Nuclei
In the discussion performed so far, the dynamics of the atomic nuclei has been assumed to be
entirely classical, and to evolve according to the Langevin Eq. (2). Such an approximation
is certainly reliable for most atomic species which are found in biomolecular systems, such as
carbon or oxygen. On the other hand, it becomes questionable for the lightest species, notably
hydrogen. Hence, in the study of reactions involving a large number of such atoms, the quantum
corrections to the stochastic dynamics of nuclei should be consistently taken into account.
To tackle this problem, we have derived the multidimensional generalization [15] of the
semiclassical extension of the Langevin Eq. (2) and of the corresponding FP equation (4), in
which quantum corrections are systematically included to order ¯h2 [13, 14]. From these equations
it is immediate to compute the quantum corrections to the DRP equations [15]. We found that
the effective action receives contribution from an additional term in the form (adopting again
for simplicity a one-dimensional notation):
where
SQ
ef f [x] =
dτ V Q
ef f [x(τ )],
(cid:90) t
0
ef f (x) (cid:39) D λ
V Q
4(kBT )3∇U (x)2∇2U (x)
is a characteristic parameter
is the correction to the effective potential and λ =
12 kBT mi
which determines the size of quantum fluctuations.
In conformational transitions driven by
the formation of hydrogen bonds, the quantum corrections to the dominant reaction pathways
have been found to give raise to sizable effects, see Ref. [15].
¯h2
(24)
(25)
Figure 4. Evolution of the end-point distance of the protein during a folding reaction, averaged
over folding trajectories obtained in DRP and MD simulations, respectively. The dashed line
represents a typical trajectory.
5. Some Applications
The DRP approach has been extensively tested and validated on toy systems [7, 8] and
simplified protein models [16, 17], and then applied to investigate realistic transitions, including
conformational transitions of peptides [6, 18], chemical reactions [12] and realistic protein folding
reactions [21]. In this section, we briefly review some of these studies.
5.1. Validation of the DRP Theory on a Coarse-Grained Model
The most straightforward way to assess the accuracy of the DRP approach consists in comparing
its predictions with those obtained directly from MD simulations. Unfortunately, for realistic
models MD simulations of protein folding reactions are not presently feasible.
Hence, we have validated the DRP approach using much simpler coarse-grained models, with
protein-like properties. For example, the model used in Ref.[17] displays a folding transition
with a well-defined and unique native state. The predictions for the time evolution of the
distance between the end-points of the chain (a typical observable in protein folding) obtained
in the DRP and MD approaches is compared in Fig. 4. We see that the biological instanton
standpoint provides a realistic description of the reactive dynamics.
5.2. Atomistically Detailed Protein Folding Simulations
The computational efficiency of the DRP approach allows to investigate a large class of reactions
which cannot be studied on the existing supercomputer using standard MD simulations. Only
very few examples of successful MD simulations of protein folding reactions have been reported
to date. These studies were either performed on a special purpose supercomputing machine in
which all the key MD algorithms are implemented at the hardware level [19] or through very long
large-scale simulations performed on several thousands of CPU's, distributed world wide [20].
On the other hand, using the DRP method [21] it was possible to simulate the same reaction
at a comparable level of statistics in less than a single day on only 32 CPU's. The DRP results
were found to be consistent with those reported in [19], and provided an explanation of the
discrepancy between the results of Ref. [19] and Ref. [20]. An illustration of one of the folding
01020Time [ps]0.51d1-16 [nm]MD (trajectory)DRP (average)MD (average)Figure 5. Folding of a peptide obtained from quantum DRP calculations. The shaded area
represents the distribution of electric charge.
trajectories obtained in the DRP approach is displayed in Fig. 1. Simulations on much larger
systems (unaccessible to MD simulations) are now being performed.
5.3. Folding of a Small Peptide from Ab-Initio Quantum Mechanical Simulations
Using the DRP approach we have performed the first study of the folding of a peptide
based on ab-initio Quantum Mechanical calculations, using the Born-Oppenheimer approach
described in section 4.3. This type of analysis could never performed using MD or Car-Parinello
approaches [3]. The results of our simulations are schematically represented in Fig. 5. This
study allowed for the first time to access the reliability of the existing classical force fields in
predicting non-equilibrium reactive trajectories. This is an important and non-trivial issue,
since the phenomenological force fields are fitted in order to reproduce quantum mechanical
calculations on static configurations, or in equilibrium condition. Hence, there is in principle no
guarantee that they should be accurate in also in the regions of phase-space which are mostly
visited by the non-equilibrium reactive trajectories. On the other hand, our studies have revealed
that classical calculations give results that are indeed in good agreement with those of quantum
ab-initio simulations. This study also allowed to quantify to what extent the electronic structure
of the chain is modified during the reaction.
6. Conclusions
Nuclear physics provides an excellent training ground where developing advanced theoretical
approaches to interacting many-particle systems.
Indeed, a number of schemes and
approximations originally proposed in this context [22] have later become standard tools in
the field of strongly correlated electrons, quantum liquids, cold atoms, . . . .
In particular, in this talk we have reviewed our recent attempt to export functional integral
techniques originally derived in QCD to investigate the non-equilibrium diffusive dynamics of
biomolecules such as proteins. We have argued that such an approach opens the door to the
study of a large number of processes which simply could not be investigated using standard MD
simulations.
Interdisciplinary studies of this type may also emphasize the importance of pushing forward
the fundamental research in nuclear theory. Indeed, they show that this activity is valuable also
in a perspective which goes beyond the traditional perimeter of nuclear physics.
!"#$%&'($)%!#!*)XiXfAcknowledgments
I would like to thank all the collaborators which have been involved in the development of the
DRP approach, and in particular H. Orland, F. Pederiva, G. Garberoglio, S. a Beccara and M.
Sega.
PF is a member of the Interdisciplinary Laboratory for Computational Science (LISC), a
joint venture of Trento University and Bruno Kessler foundation. The numerical simulations
were performed on the Aurora supercomputer at LISC. This research was partially funded by
the Provincia Autonoma di Trento and by INFN, through the AuroraScience project.
References
[1] P. L. Freddolino et al., Biophys. J. 94, L75 (2008).
[2] A.V. Finkelstein and O.B. Ptitsyn, "Protein Physics: a Course of Lectures", Academic Press, London (2002).
[3] A.R. Leach, "Molecular modeling: principle and applications" (2nd ed.) Pearson Education (Harlow,
England), 2001.
[4] F. Schwabl, "Statistical Mechanics" (2nd ed.), Springer-Verlag, Berlin 2006.
[5] P. Faccioli, M. Sega, F. Pederiva and H. Orland, Phys. Rev. Lett. 97, 108101 (2006).
[6] M. Sega, P. Faccioli, F. Pederiva, G. Garberoglio and H. Orland, Phys. Rev. Lett. 99, 118102 (2007).
[7] E. Autieri, P. Faccioli, M. Sega, F. Pederiva and H. Orland, J. Chem Phys. 130, 064106 (2009).
[8] G. Mazzola, S. a Beccara, P.Faccioli, and H. Orland, J. Chem. Phys. 134, 164109 (2011).
[9] R. Elber, and D. Shalloway, J. Chem. Phys. 112, 5539 (2000).
[10] O. Corradini, P. Faccioli and H. Orland, Phys. Rev. E80 061112 (2009) .
[11] P. Faccioli, J. Chem. Phys. 133 164106 (2010).
[12] S. a Beccara, G. Garberoglio, P. Faccioli and F. Pederiva, J. Chem. Phys. 132, 111102 (2010).
[13] J. Ankerhold, P. Pechukas and H. Grabert, Phys. Rev. Lett. 87, 086802 (2001). J. Ankerhold and H. Grabert,
Phys. Rev. Lett. 101, 119903 (2008) (Erratum). J. Ankerhold, Phys. Rev. E64, 060102 (2001).
[14] W. T. Coffey, Y. P. Kalmykov, S. V. Titov, and B. P. Mulligan, J. Phys. A 40, F91(2007). W. T. Coffey, Y.
P. Kalmykov, S. V. Titov, AND L. Cleary, Phys. Rev. E 78 031114 (2008).
[15] S. a Beccara, G. Garberoglio and P. Faccioli, J. Chem. Phys. 135 034103 (2011).
[16] P. Faccioli, J. Phys. Chem. B112, 137560 (2008).
[17] P. Faccioli, A. Lonardi and H. Orland, J. Chem. Phys. 133, 045104 (2010).
[18] S. a Beccara, P. Faccioli, M. Sega, G. Garberoglio, F. Pederiva and H. Orland, J. Chem. Phys. 134, 024501
(2011).
[19] D.E. Shaw et al., Science 330, 341 (2010).
[20] D.E. Ensign and V.J. Pande, Biophys. Journ. 96 L53 (2009).
[21] S. a Beccara, T. Skrbic, R. Covino and P. Faccioli, presently under review.
[22] J.W. Negele and H. Orland, "Quantum many-particle systems", Westview Press 1998.
|
1906.09717 | 1 | 1906 | 2019-06-24T04:26:10 | Emergence of localized persistent weakly-evanescent cortical brain wave loops | [
"physics.bio-ph",
"nlin.PS"
] | An inhomogeneous anisotropic physical model of the brain cortex is presented that predicts the emergence of non--evanescent (weakly damped) wave--like modes propagating in the thin cortex layers transverse to both the mean neural fiber direction and to the cortex spatial gradient. Although the amplitude of these modes stays below the typically observed axon spiking potential, the lifetime of these modes may significantly exceed the spiking potential inverse decay constant. Full brain numerical simulations based on parameters extracted from diffusion and structural MRI confirm the existence and extended duration of these wave modes. Contrary to the commonly agreed paradigm that the neural fibers determine the pathways for signal propagation in the brain, the signal propagation due to the cortex wave modes in the highly folded areas will exhibit no apparent correlation with the fiber directions. The results are consistent with numerous recent experimental animal and human brain studies demonstrating the existence electrostatic field activity in the form of traveling waves (including studies where neuronal connections were severed) and with wave loop induced peaks observed in EEG spectra. The localization and persistence of these cortical wave modes has significant implications in particular for neuroimaging methods that detect electromagnetic physiological activity, such as EEG and MEG, and for the understanding of brain activity in general, including mechanisms of memory. | physics.bio-ph | physics | Emergence of localized persistent weakly -- evanescent cortical brain wave loops
Vitaly L. Galinsky∗
Center for Scientific Computation in Imaging, University of California at San Diego, La Jolla, CA 92037-0854, USA and
Department of ECE, University of California, San Diego, La Jolla, CA 92093-0407, USA
Center for Scientific Computation in Imaging, University of California at San Diego, La Jolla, CA 92037-0854, USA and
Center for Functional MRI, University of California at San Diego, La Jolla, CA 92037-0677, USA
(Dated: June 25, 2019)
Lawrence R. Frank†
An inhomogeneous anisotropic physical model of the brain cortex is presented that predicts the
emergence of non -- evanescent (weakly damped) wave -- like modes propagating in the thin cortex lay-
ers transverse to both the mean neural fiber direction and to the cortex spatial gradient. Although
the amplitude of these modes stays below the typically observed axon spiking potential, the lifetime
of these modes may significantly exceed the spiking potential inverse decay constant. Full brain nu-
merical simulations based on parameters extracted from diffusion and structural MRI confirm the
existence and extended duration of these wave modes. Contrary to the commonly agreed paradigm
that the neural fibers determine the pathways for signal propagation in the brain, the signal propa-
gation due to the cortex wave modes in the highly folded areas will exhibit no apparent correlation
with the fiber directions. The results are consistent with numerous recent experimental animal and
human brain studies demonstrating the existence electrostatic field activity in the form of traveling
waves (including studies where neuronal connections were severed) and with wave loop induced
peaks observed in EEG spectra. The localization and persistence of these cortical wave modes has
significant implications in particular for neuroimaging methods that detect electromagnetic phys-
iological activity, such as EEG and MEG, and for the understanding of brain activity in general,
including mechanisms of memory.
The majority of approaches to characterizing brain dy-
namical behavior are based on the assumption that sig-
nal propagation along well known anatomically defined
pathways, such as major neural fiber bundles, tracts or
groups of axons (down to a single axon connectivity)
should be sufficient to deduce the dynamical character-
istics of brain activity at different spatiotemporal scales.
Experimentally, data on which this assumption is em-
ployed range from high temporal resolution neural oscil-
lations detected at low spatial resolution by EEG/MEG
[1] to high spatial resolution functional MRI resting state
modes oscillation detected at low temporal resolution
[2]. As a consequence, a great deal of research activ-
ity is directed at the construction of connectivity maps
between different brain regions (e.g., the Human Connec-
tome Project [3]), and using those maps to study dynami-
cal network properties with the help of different models of
signal communication through this network along struc-
turally aligned pathways [4], i.e. by allowing input from
different scales or introducing axon propagation delays
[5].
However, recent detection [6] of cortical wave activity
spatiotemporally organized into circular wave-like pat-
terns on the cortical surface, spanning the area not di-
rectly related to any of the structurally aligned pathways,
but nevertheless persistent over hours of sleep with mil-
lisecond temporal precision, presents a formidable chal-
lenge for network theories to explain such a remarkable
∗ [email protected]
† [email protected]
synchronization across a multitude of different local net-
works. Additionally, many studies show evidence that
electrostatic field activity in animal or human hippocam-
pus (as well as cortex) are traveling waves [7] that can
affect neuronal activity by modulating the firing rates [8]
and may possibly play an important functional role in
diverse brain structures [9]. And perhaps more impor-
tantly it has been experimentally shown [10] that peri-
odic activity can self-propagate by endogenous electric
fields even through a physical cut in vitro that destroys
all mechanisms of neuron to neuron communication.
In this paper we investigate a more general physical
wave mechanism that allows cortical surface wave prop-
agation in the cross fiber directions due to the interplay
between tissue inhomogeneity and anisotropy in the thin
surface cortex layer. This new mechanism has been over-
looked by previous models of brain wave characterization
and thus is absent from current network pathway recon-
struction and analysis approaches.
The main claim of this paper is that there is a sim-
ple and elegant physical mechanism behind the existence
of these cross-fiber waves that can explain the emer-
gence and persistence of wave loops and wave propaga-
tion along the highly folded cortical regions with a rel-
atively slow damping. The lifetime of these wave -- like
cortex activity events can significantly exceed the decay
time of the typical axon action potential spikes and thus
can provide "memory -- like" response in the cortical areas
generated as a result of "along -- the -- axon" spiky activa-
tions. This new mechanism may provide an alternative
approach for the integration of microscopic brain prop-
erties and for the development of a "physical" model for
9
1
0
2
n
u
J
4
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
1
7
9
0
.
6
0
9
1
:
v
i
X
r
a
memory.
The paper derives cortex wave dispersion relation from
well known relatively basic physical principles and pro-
vides illustrations of why and how cortical tissue inho-
mogeneity and anisotropy influence propagation magni-
tude, time-scales, and directions and supports extended
and highly structured regions of existence in dissipative
media using simple 1- and 2-D anisotropic models, as
well as more realistic full brain model based on a set
of parameters extracted from real diffusion and struc-
tural MRI acquisitions. The wave properties (frequency
ranges, phase and group velocities, possible spectra) are
then compared with real EEG wave acquisitions. Finally,
examples of wave propagation are studied analytically
and numerically using a simple idealized but informa-
tive spherical shell cortex model (i.e. thin inhomogeneous
layer around a sphere with homogeneous anisotropic con-
ducting medium) as well as a more realistic anisotropic
inhomogeneous full brain model with actual cortical fold
geometry that clearly shows the emergence of localized
persistent wave loops or rotating wave patterns at vari-
ous scales, including at similar scales of global rotation
recently detected experimentally [6].
An important aspect of the cortical waves model we
present is that it is based on relatively simple but phys-
ically motivated averaged electrostatic properties of hu-
man neuronal tissue within realistic data-derived brain
tissue distributions, geometries, and anisotropy. While
the source of these averaged tissue properties includes
the extraordinarily complex network of neuronal fiber
connections supporting a multitude of underlying cellu-
lar, subcellular and extracellular processes, we demon-
strate that the inclusion of such details for the activa-
tion/excitation process is not necessary to produce co-
herent, stable, macroscopic cortical waves.
Instead, a
simple and elegant physical model for wave propagation
in a thin dissipative inhomogeneous and anisotropic cor-
tical layer of these averaged properties is sufficient to pre-
dict the emergence of coherent, localized, and persistent
wave loop patterns in the brain.
We will start with the most general form of description
of brain electromagnetic activity using Maxwell equa-
tions in a medium
∂ρ
∂t
∂D
∂t ⇒
∇ · D = ρ, ∇ × H = J +
+ ∇ · J = 0.
Using the electrostatic potential E = −∇φ, Ohm's law
J = σ · E (where σ ≡ {σij} is an anisotropic conductiv-
ity tensor), a linear electrostatic property for brain tis-
sue D = εE, assuming that the permittivity is a "good"
function (i.e.
it does not go to zero or infinity every-
where) and taking the change of variables ∂x → ε∂x(cid:48),
the charge continuity equation for the spatial-temporal
evolution of the potential φ can be written in terms of a
permittivity scaled conductivity tensor Σ = {σij/ε} as
(cid:0)
∇2φ(cid:1) = −∇ · Σ · ∇φ + F,
∂
∂t
(1)
2
where we have included a possible external source (or
forcing) term F. For brain fiber tissues the conductivity
tensor Σ might have significantly larger values along the
fiber direction than across them. The charge continuity
without forcing i.e., (F = 0) can be written in tensor
notation as
∂t∂2
i φ + Σij∂i∂jφ + (∂iΣij) (∂jφ) = 0,
where repeating indices denotes summation.
Sim-
ple linear wave analysis,
i.e. substitution of φ ∼
exp (−i(k · r − Ωt)), gives the following complex disper-
sion relation
(2)
D(Ω, k) = −iΩk2
i − Σijkikj − i∂iΣijkj = 0,
(3)
which is composed of the real and imaginary components:
γ ≡ (cid:61)Ω = Σij
kikj
k2
ω ≡ (cid:60)Ω = −
∂iΣijkj
k2
(4)
The condition for non -- (or weak -- ) evanescence is that the
oscillatory (i.e., imaginary) component of φ, character-
ized by the frequency ω, is much larger than the decaying
(i.e., real) component, characterized by the damping γ:
i.e. that the condition γ/ω (cid:28) 1 must be satisfied. This
requirement is clearly not satisfied if reported average
isotropic and homogeneous parameters are used to de-
scribe brain tissues. For typical low frequency ((cid:46) 10Hz)
white and gray matter conductivity and permittivity (i.e.
from [11]) εGM = 4.07 · 107ε0, εW M = 2.76 · 107ε0,
σGM = 2.75 · 10−2 S/m, σW M = 2.77 · 10−2 S/m, where
ε0 = 8.854187817·10−12 F/m is the vacuum permittivity)
the damping rate γ is in the range of 75 -- 115 s-1 which
would be expected to give strong wave damping.
In order to better understand the effects of brain tis-
sue micro- and macro-structure on the manifestation of
propagating brain waves, it is instructive to consider two
idealized tissue models. In the first model (Fig. 1a) all
brain fibers are packed in a half space aligned in z di-
rection and their number decreases in x direction in a
relatively thin layer at the boundary. We assume that
small cross fiber currents can be characterized by a small
parameter and introduce the conductivity tensor as
υ υ υ
υ υ 0
υ 0 υ
.
Σ =
(5)
where υ ≡ υ(x). For the υ(x) dependence we will as-
sume that the conductivity is changing only through a
relatively narrow layer at the boundary (as illustrated in
Fig. 1a) and the conductivity gradient is directed along
x axis. Then we will look for a solution for the poten-
tial φ located in the thin layer of inhomogeneity (that is
we substitute x → x) that depends on z and y only as
φ = φ(cid:107)(z) + φ⊥(y).
0 :
1 :
∂
∂t
∂
∂t
∂2φ(cid:107)
∂z2 + υ
∂2φ⊥
∂y2 +
∂2φ(cid:107)
∂z2 +
∂φ⊥
∂υ
∂x
∂y
∂υ
∂x
∂φ(cid:107)
∂z
= 0
+ O() = 0
(6)
(7)
3
In order to extend the above analysis to a more re-
alistic model of brain tissue architecture we extracted
volumetric structural brain parameters from high reso-
lution anatomical MRI datasets as well as brain fiber
anisotropy from diffusion weighted MRI datasets. All
anatomical and diffusion MRI datasets are from the Hu-
man Connectome Project [3]. The details for all of the
processing steps can be found in [14 -- 16]. More refined
procedures for constructing the conductivity tensor and
anisotropy in different brain regions, cortical areas in par-
ticular, would clearly be beneficial and will be addressed
in the future.
To provide an illustration of where the conductivity
anisotropy and inhomogeneity can form appropriate con-
ditions for cortex surface waves generation we created
plots that can be used to characterize the ratio γ/ω.
To construct the ratio we calculated two vectors, ∂iΣij
and Σijki and compared their norms. For wave vector
k we used a vector with the same direction as in ∂iΣij
vector and magnitude k = ∇Σ/Σ, i.e. our intention
is to compare norms of dissipative and wave-like terms
at kh ≈ 1 where h is the cortical thickness. Fig. 2 shows
plot of the ratio of Σijki and ∂iΣij at two different
depths inside the brain with dissipative regime in the in-
ner cortex (left) and wave -- like conditions in the outer
cortex (right).
FIG. 2.
Isovolume maps for comparison of dissipation vs
wave -- like effects at different cortex layers. The left panel
shows the isosurface from the cortex area that may be rep-
resentative of white -- gray matter interface. The right panel
shows the isosurface that is located in the outer cortex area
of gray matter. The color scheme uses shades of green to
mark regions where dissipative term dominates, i.e. Σijki ≥
∂iΣij, shades of red where Σijki < ∂iΣij ≤ 2Σijki, and
shades of blue where the wave-like term is more than two
times dominant, i.e. 2Σijki < ∂iΣij. The inner cortex
shown in the left panel clearly display prevalence of dissipa-
tion, whereas the outer cortex shown in the right panel allows
for wave -- like cortex activity in a majority of the locations.
Both anisotropy and inhomogeneity are important for
the existence of the cortex surface waves. For example
for inhomogeneous but isotropic tissue the conductivity
tensor Σij will simply be Sδij, where S is a scalar inho-
mogeneous conductivity. Therefore both the phase ve-
locity (vph = ω/k) and the group velocity (vgr = ∂ω/∂k)
will include terms ∇S and ∇S · k, meaning that those
waves are not able to propagate normally to the local
FIG. 1. (a) Schematic picture of half-plane packing of fibers.
The uniform area of fibers oriented along z direction (shown
in green) is bounded by a thin transitional area (magenta)
where the conductivity gradient may be important (a sketch
of one possible conductivity profile is shown in the bottom
panel). (b) Schematic picture that can be used as a crude
two dimensional approximation of fold. The direction of fiber
conductivity has only x and z components and all quantities
are assumed to be uniform in y direction.
where 0 and 1 denote the zeroth and the first orders of
power.
The first equation (6) describes a potential along the
fiber direction and is a damped oscillator equation that
has a decaying solution. But the second equation (7)
describes a potential perpendicular to the fiber direction
and does not include a damping term, hence it describes a
pure wave-like solution that propagates in the thin layer
transverse to the main fiber direction. Thus although
this wave-like solution φ⊥ has a smaller amplitude than
along the fiber action potential φ(cid:107), it can nevertheless
have a much longer lifetime.
We would like to stress that the equations (6) and (7)
are here only for illustrative purposes to reiterate a rel-
atively obvious but often overlooked consideration that
under anisotropic inhomogeneous conditions some direc-
tion may happen to be better suited for wave propagation
than the others. Those equations should not be consid-
ered as the most important part as the complete disper-
sion relation (3) will be used to study waves propagation
later.
In order to account for geometric variations, we con-
struct a slightly more complex two dimensional model
that can be viewed as a very crude approximation of a
cortical fold (Fig. 1b). The equation for the φ⊥(y) will
again be of a wave type, similar to (7), with the addition
of a z component of the conductivity gradient and with a
similar wave -- like solution that will include an additional
term induced by the inhomogeneity in z.
The sole purpose of those examples is to provide illus-
trations that dissipative media with complex structure
may show surface wave -- like solutions. Surface waves at
the boundary of various elastic media have been exten-
sively studied and used in various areas of science (in-
cluding acoustics, hydrodynamics, plasma physics, etc.)
since the work of Lord Rayleigh [12]. The existence of
surface waves at the dissipative medium boundary is also
known [13].
zxy·····················υxzxyab00.511.522.533.54i ij/ij ki∂iΣij/Σijkiconductivity gradient ∇S. This restriction is absent in
cortex areas when both anisotropy and inhomogeneity
are present.
To provide some (possibly overly optimistic) estimates
based on a typical human brain dimensions we can con-
sider a simple spherical cortex shell model with a cortical
layer of fixed thickness h ≈ 1.5 -- 3mm spread over a hemi-
sphere of radius R ≈ 75mm, with all parameters kept
constant inside the hemisphere (for r < R) and changing
as a function of radius r in a cortical layer. Even without
taking into account the known strong anisotropy of neu-
ral tissue, these simple geometric considerations provides
for the longest waves (with the smallest amount of damp-
ing) with γ/ω ∼ kh ∼ 0.02 -- 0.04. Anisotropy will reduce
this estimate even further, thus further strengthening the
condition necessary to support stable waves. This simple
spherical shell model can be used to illustrate the nat-
ural appearance of standing -- type cortical waves, which
can be easily understood from simple geometrical optics
arguments. Using the dispersion relation (3) with only
single component of the conductivity tensor Σzz ≡ S(r),
the wave frequency ω can be expressed as
ω = −
1
r
dS
dr
kzz
k2 ;
(8)
where we have neglected the term Sk2
Then from the geometrical optics ray equations
z because kh (cid:28) 1.
drl
dt
=
∂ω
∂kl
,
dkl
dt
∂ω
∂rl
,
= −
we can get
dx
dt
dy
dt
dz
dt
= −2ω
= −2ω
= −ω
kx
k2 ,
ky
k2 ,
z − k2
2k2
k2kz
,
dkx
dt
dky
dt
dkz
dt
= −
= −
= −
dω
dr
dω
dr
1
z
,
x
r
y
r
,
(cid:20) dω
dr
(cid:21)
+ ω
.
z2
r
(9)
(10)
z =(cid:112)
These equations will generate rays inside the spheri-
cal cortex shell showing wave propagation across both
the fibers and the conductivity gradient in the cortex
subregion where ω dω/dr < 0. For kz = k/√2 and
−ωr/(dω/dr) the wave path has the simplest form
-- the wave follows the same loop through the cortex over
and over again. Different families of frequencies ω and
wavevectors k will result in the appearance of cortical
wave loops at different cortex locations.
In order to determine a possible energy distribution
across different frequencies for these cortical waves some
knowledge about the forcing term F in (1) is required.
A rough estimate for this distribution in the form of a
power spectra scaling can be carried out using some sim-
ple assumptions. Assuming that the forcing consists of
spiking input localized at random locations and times,
it can be described as a sum of delta functions, F =
i Aiδ(t − ti)δ(r − ri), corresponding to a flat forcing
(cid:80)
4
FIG. 3. Spectral power of EEG signal collected with 64 sen-
sor array and averaged over all sensors for six independent
subjects is shown in six panels. The dashed lines outline the
−2 in the lower (f (cid:46) 1.2Hz) and higher (f (cid:38) 92Hz)
predicted f
parts of the spectra. The dashed -- dotted vertical lines denote
the frequency range where the cortical wave loops may be
generated. Both the slope and the range agree very well with
typically observed values.
the last term in (2) (∂iΣij)(∂jφ) ∼ ω(cid:112)
frequency spectrum in the Fourier domain. Then, from
φω2 ∼ const,
we can estimate the exponent α = 2 in a power law
scaling of the potential φ frequency spectrum (i.e. for
φω2 ∼ ω−α).
The presence of the cortical wave loops described above
can modify this ω−2 dependence and thus modify the
spectrum in such a way that spectrum peaks are gener-
ated that correspond to these loop wave currents. From
dS/dr ∼ Σzz/h, z =(cid:112)
(8) we can estimate the range of frequencies where those
loops can possibly be present. Taking 1/r ∼ 1/R,
−ωr/(dω/dr) ∼ √Rh gives a fre-
quency estimate as f = ω/2π ∼ Σzz/(2πk√2Rh), hence
for the largest and smallest wavenumbers defined by the
smallest (1.5mm) and largest (75mm) loop radii, the fre-
quency f spans the range 1.2 -- 92Hz (vph ≈ 0.002−7m/s).
The large scale circular cortical waves of [6] (9 -- 18Hz, 2 --
5m/s) are clearly inside this range. The above spherical
shell wave propagation example (as well as wave simula-
tions presented below) are not able to predict the exact
values for wave amplitude as it requires calculation of the
balance between excitation and dissipation of these waves
at various frequency ranges and the simple estimates us-
ing amplitudes of spiking are not that particularly infor-
mative as they simply predict the values that are in the
range of typically observed field potentials. Neverthe-
less, even without the amplitude estimation our analysis
is able to predict the preferred directions of wave propa-
gation and loop pattern formation based on geometrical
and tissue properties. Moreover, this framework provides
the mechanisms to incorporate a more complete quanti-
tative description of axonal spiking, which is beyond the
scope of this paper but currently under investigation in
10-210010210-1010-510010-210010210-1010-510010-210010210-1010-510010-210010210-1010-510010-210010210-1010-510010-210010210-1010-5100Spectralpowerφf2(arbitraryunits)Frequencyf(Hz)Frequencyf(Hz)our lab.
Evidence for the existence of these wave loop induced
spectral peaks is shown in Fig.3, which shows the spectral
power of the EEG signal for six subjects [17], averaged
over all sensors. The dashed lines outline the predicted
f−2 in the lower (f (cid:46) 1.2Hz) and higher (f (cid:38) 92Hz)
parts of the spectra. The dashed -- dotted vertical lines
denote the frequency range where the cortical loops may
exist, agreeing very well with typically observed EEG
excessive activity range, from low frequency delta (0.5 --
4Hz) to high frequency gamma (25 -- 100Hz) bands.
The effect described theoretically above can be demon-
strated through numerical simulation of wave propaga-
tion in a thin dissipative inhomogeneous and anisotropic
cortical layer. As a starting point for numerical study
we included Fig. 4, that shows spatial snapshots of
the dynamical behavior of randomly generated wave
trajectories (the movies are in [18]) in a regime with
γeff Lloop/vgr ≤ 1 (where γeff is an effective wave dissi-
pation rate, i.e. a difference between average dissipation
and spiky activations rates for a wave packet propagating
with group velocity vgr along some characteristic loop of
Lloop length. Wave packets are simulated using the ray
equations (9) where the general form of anisotropic dis-
persion relation (3) and (4) is used.
In the idealized spherical model some of the emer-
gent persistent localized cortical wave patterns are pre-
cisely the simple loop pattern predicted by (10), despite
the complex initial spatiotemporal pattern of the initial
wave trajectory. As further predicted, the simulations
informed by the real human data with the same cortex
fold geometry as in Fig. 2 (middle and right panels) also
produce stable loop structures, now embedded within the
complex geometry of the cortical folds.
All wave simulations were initialized with wave pack-
ets of random parameters (frequency, wave number, lo-
cation, etc), but clearly show an emergence of localized
persistent closed loop patterns at different spatial and
temporal scales, from scales as large as global the whole
brain rotational wave activity experimentally detected in
[6] to as small as the resolution used for the cortical layer
thickness detection.
In conclusion, in this paper we have presented an inho-
mogeneous anisotropic physical model of wave propaga-
tion in the brain cortex. The model predicts that in ad-
dition to the well-known damped oscillator -- like wave ac-
tivity in brain fibers, there is another class of brain waves
that are not directly related to major fibers, but instead
propagate perpendicular to the fibers along the highly
folded cortical regions in a weakly -- evanescent manner
that results in their persistence on time scales long com-
pared to waves along brain fibers. Thus waves can poten-
tially propagate in any direction, including the direction
along the fibers or in the direction of the inhomogene-
ity gradient. However, the dissipation of the waves is
smallest when they propagate cross fibers, therefore, on
average the cross fiber direction of propagation should be
seen more often. For the first time, we have obtained the
5
FIG. 4. Complete wave packet trajectory snapshots (upper
row) and emergent stable wave loop patterns (lower row) for
the thin spherical shell cortex model (left) and for the realis-
tic cortex fold geometry (middle and right). All wave packets
were initialized with random parameters assuming the pres-
ence of spiky activation sources (not shown). Movie files for
these and additional loop examples can be found in [18].
dispersion relation for those surface cortex waves, and
have shown, both analytically and numerically, a plau-
sible argument for their existence. Through numerical
analysis we have developed a procedure to generate an
inhomogeneous and anisotropic distribution of conduc-
tivity tensors using anatomical and diffusion brain MRI
data. While the detailed numerical studies of effects and
importance of these waves and their possible biological
role are out of the scope of this paper, we have presented
preliminary results that suggest that the time of life for
these wave -- like cortex activity events may significantly
exceed the decay time of the typical axon action poten-
tial spikes. Thus, they can provide an persistent neu-
ronal response in the cortical areas generated as a result
of "along -- the -- axon" spiky activations.
The interaction of the traveling waves with spiking ac-
tivity is of course an interesting and important question.
Our model supports the inclusion of spiking activity as a
source term to the wave model (one example is sketched
in appendix D. In particular this may allow the devel-
opment of models of brain rhythm generation based on
coupling with spiking sources. Exploration of the inter-
play of the traveling waves and spiking activity with this
model will certainly be the focus of future work.
The ranges of parameters for the waves produced by
our model are in agreement with presented by several
studies [7] that present evidence that electrostatic field
activity in several areas of animal and human brains
are traveling waves that can affect neuronal activity by
modulating the firing rates [8] and may possibly play
an important functional role in diverse brain structures
[9]. The natural self organization of these traveling
waves into loop -- like structures that our model produces
agrees well with recently detected [6] cortical wave ac-
CompletewavetrajectoryEmergentlooppatterntivity spatiotemporally organized into circular wave-like
patterns on the cortical surface. Self-propagation of en-
dogenous electric fields through a physical cut in vitro
when all mechanisms of neuron to neuron communication
has been destroyed [10] can potentially be attributed to
these waves as well. We have also demonstrated that the
peaks these wave loops would induce in EEG spectra are
consistent with those typically observed EEG data.
Direct experimental results have shown that even de-
spite the small amplitudes of the external field potentials
relative to the threshold of a spiking neuron, external
fields can play a substantial role in the spiking activ-
ity and "even very small and slowly changing fields that
triggered Ve changes under 0.2 mV led to phase locking
of spikes to the external field and to a greatly enhanced
spike-field synchrony" [19]. Therefore our models ability
to predict regions of cortex where external wave activity
can emerge and form a sustained loop pattern has the
potential to be important for understanding where the
neuron spiking synchronization will have better chances
to be achieved, hence it can provide substantial input in
understanding effects on neural information processing
and plasticity.
The ability of this new physical model for the gen-
eration, propagation, and maintenance of brain waves
may have significant implications for the analysis of elec-
trophysiological brain recording and for current theories
about human brain function. Furthermore, the depen-
dence of these waves on brain geometry, such as corti-
cal thickness, has potentially significant implications for
understanding brain function in abnormal states, such
6
as Alzheimer's Disease, where cortical thickness changes
are evident, and the dependence on tissue status may
be important in conditions such as Traumatic Brain In-
jury, where tissue damage may alter its anisotropic and
inhomogeneous properties.
ACKNOWLEDGEMENTS
LRF and VLG were supported by NSF grants DBI-
1143389, DBI-1147260, EF-0850369, PHY-1201238, ACI-
1440412, ACI-1550405 and NIH grant R01 MH096100.
Data were provided [in part] by the Human Connectome
Project, WU-Minn Consortium (Principal Investigators:
David Van Essen and Kamil Ugurbil; 1U54MH091657)
funded by the 16 NIH Institutes and Centers that sup-
port the NIH Blueprint for Neuroscience Research; and
by the McDonnell Center for Systems Neuroscience at
Washington University. Data collection and sharing for
this project was provided by the Human Connectome
Project (HCP; Principal Investigators: Bruce Rosen,
M.D., Ph.D., Arthur W. Toga, Ph.D., Van J. Weeden,
MD). HCP funding was provided by the National Insti-
tute of Dental and Craniofacial Research (NIDCR), the
National Institute of Mental Health (NIMH), and the
National Institute of Neurological Disorders and Stroke
(NINDS). HCP data are disseminated by the Laboratory
of Neuro Imaging at the University of Southern Califor-
nia. The EEG data is courtesy of Antigona Martinez of
the Nathan S. Kline Institute for Psychiatric Research.
Appendix A: Spherical shell cortex model
We first provide details about the procedures used for generating inhomogeneous and anisotropic components of
the permittivity scaled conductivity tensor Σ.
Spherical shell cortex model is represented by fixed anisotropy tensor σij(r) ≡ σij scaled by a radial inhomogeneous
density ρ(r) ≡ ρ(r), such that the total conductivity tensor Σij(r) is defined by
where σij is a constant, anisotropic, positive semidefinite symmetric tensor and ρ(r) is piecewise continuous function
of radius r (0 ≤ r ≤ 1)
Σij(r) = ρ(r)σij(r),
(A1)
ρ(r) =
1
0
(cid:18)
(cid:18)
1 + arctan
α∞
r − r0
r1 − r0
2 arctan (α∞)2n+1
1 − 2
(cid:19)(cid:19)2n+1
r ≤ r0
r0 < r < r1
r ≥ r1
The spherical shell cortex wave simulations include three different anisotropic tensor σij choices,
ε 0 0
0 ε 0
0 0 1
σ(1) =
1 + 2ε 1 − ε
1 − ε
1 − ε 1 + 2ε 1 − ε
1 − ε 1 + 2ε
1 − ε
σ(2) =
1
3
1 0 0
0 ε 0
0 0 1
σ(3) =
7
(A2)
where (for ε = 0) σ(1) represents the conductivity tensor used in analytical solution of equations (10) of the paper,
i.e. currents only in z direction, σ(2) represents different orientation, where currents are allowed in the direction of
45◦ relative to all axis, and σ(3) represents more complicated current anisotropy, with crossing currents flowing in x
and z direction, but with no currents in y direction.
Parameters n, r0 and r1 used to control the thickness of the inhomogeneous "cortical" layer (r0=0.5, r1=0.9,
α∞=500 and n =0,2,14 and 25 used for Figs. A1 to A4).
Appendix B: Cortical fold model
For cortical fold model the conductivity tensor Σij(r) is again defined as a product of inhomogeneous ρ(r) and
anisotropic σij(r) parts (A1), but both parts are now functions of brain locations.
1.
Inhomogeneity estimation
The inhomogeneous density function ρ(r) is estimated from high resolution anatomical MRI data by processing
it with SWD [14] (skull stripping, field of view normalization, noise filtering) and then registering to MNI152 space
[20, 21] with SYM-REG [16]. The final 1mm3 182x218x182 volume is used as the inhomogeneous density ρ(r).
For the anisotropy tensor σij(r) several different test cases were employed.
2. Anisotropy estimation
a. Fixed anisotropy orientation and value
The same σ(1), σ(2) and σ(3) fixed (location independent) anisotropic tensors as in the spherical shell cortex model
(A2) were assigned to every location in the brain. This is the simplest case that may allow to separate the effects of
inhomogeneity and anisotropy on loop pattern formation.
b. Varying anisotropy orientation and fixed anisotropy value
Multiple diffusion direction and diffusion gradient strength MRI data were used to estimate diffusion tensor Dij
[22].
The anisotropy orientation was estimated using eigenvector d(1) of the diffusion tensor Dij with the largest eigen-
value λ(1)
d . The anisotropy tensor σij(r) was defined as
σ(r) = RT σ(1)R
(B1)
where the value of anisotropy is constant across the volume (ε = 0, 0.01 and 0.1 were used for different test examples).
R is a rotation matrix between directions (0,0,1) and d(1) ≡ (x, y, z) that can be expressed as
1 − x2
1+z
− xy
−x
R =
1+z
1+z − xy
1 − y2
−y
1+z
x
y
1 − x2+y2
1+z
(B2)
c. Varying anisotropy orientation and value
8
Similarly to the previous case the diffusion tensor Dij was estimated for each voxel and eigenvector d(1) with
d was used to define the anisotropy tensor major axis. The position dependent anisotropy
(in both cases resulting in
d and λ(3)
for σ22 and
the largest eigenvalue λ(1)
value was defined by intoducing parameter ε as either λ(2)
axisymmetric form of the conductivity tensor) and also using two different values λ(2)
σ11.
d )/2/λ(1)
d + λ(3)
d /λ(1)
d /λ(1)
d /λ(1)
d
or (λ(2)
d
d
d. Varying anisotropy orientation and value estimated from microstructure fiber anisotropy
The microstructure anisotropy at the level of a single cell and its extracellular vicinity may be significantly higher
than detected by diffusion MRI estimates. Recent measurements of effective impedance that involve both intercellular
and extracellular electrodes [23] show significantly lower conductivity (by orders of magnitude) than conductivity
obtained by measurements between extracellular electrodes only. Although attributing this to lower than typically
assumed conductivity (or higher impedance) of extracellular medium may be questionable [24], this clearly confirms
high anisotropy for the effective conductivity that should be used for analysis of effects of intercellular and membrane
sources (and these are the most important type of sources for generation of surface brain waves considered in this
paper). This may be important for future generation and refinement of the macroscopic conductivity estimates from
microstructure data.
We assumed here that anisotropic form of σ(1)
ij with ε = 0.001, 0.01 or 0.1 can be used to describe different levels
of intercellular conductivity as well as microstructure extracellular conductivity in the vicinity of cell membranes
(where z-direction corresponds to the direction of the fiber). Using full brain tractography results [25] we generated
the anisotropic part of the conductivity tensor σ(r) in voxel at r location as an average over all fiber orientations
assuming N fibers inside the voxel with Rk orientation matrix for every fiber k, i.e.
N(cid:88)
k
σ(r) =
1
N
RT
k σ(1)Rk,
(B3)
with the complete form of the conductivity tensor Σ again given by (A1).
Appendix C: Wave trajectory integration
To obtain the trajectory of a brain wave with frequency ω emitted at a point r0 with an initial wave vector k0 we
substitute the inhomogeneous anisotropic conductivity tensor Σ in the wave dispersion relation (imaginary part of
eq. (3) of the paper)
and obtain wave ray equations as
D(ω, k) = ωk2 + ∂iΣijkj = 0,
(cid:90)
0
k2dτ
=
= k2,
t =
= 0,
ω = const
dt
dτ
dω
dτ
drl
dτ
dkl
dτ
∂D(ω, k)
∂ω
∂D(ω, k)
∂t
∂D(ω, k)
∂kl
∂D(ω, k)
∂rl
= −
=
= −
⇒
⇒
= 2ωkl + ∂iΣil,
= −∂l∂iΣijkj.
(C1)
(C2)
(C3)
(C4)
(C5)
We integrate these equations starting at τ =0, t = 0, r = r0 and k = k0 to trace the wave trajectory t(τ ), r(τ ) and
k(τ ).
For numerical integration it is beneficial to split k into magnitude k and direction k parts (k = kk, and k2 = 1),
and rewrite the last two equations as
drl
dτ
dkl
dτ
= −2∂iΣij
kj
kl + ∂iΣil,
= −∂l∂iΣij
kj + ∂m∂iΣij
kj
km
kl,
9
(C6)
(C7)
where in the last equation the right hand side is orthogonal to k to guarantee that k2 = 1, and the algebraic
expression (C1) was substituted for the differential equation for wave vector magnitude k.
Appendix D: Wave dissipation and excitation
An integral along the wave trajectory t(τ ), r(τ ) and k(τ )
(cid:90)
(cid:90)
0
−
−
0
W = W0 exp
= W0 exp
[γdis (r(τ )) − γexc (r(τ ))] dt
(cid:104)
Σij (r(τ )) ki(τ )kj(τ ) − γexc (r(τ ))
(cid:105)
k2(τ )dτ
(D1)
with any appropriate model form of wave excitation γexc describes a change of wave energy W along its path. This
may allow the study of many interesting questions of brain wave dynamics, i.e. to identify potential active area where
wave intensity may grow as a result of certain frequency and/or spatial distributions of neuronal spiky activation, or
to estimate the spatial extent of coherent activation area as a result of some particular point sources, or to find when
and/or how these waves may potentially become important in triggering neuronal firing, that is to study possible
mechanisms of synchronization and feedback, etc.
Appendix E: Persistent wave loop patterns
One of the interesting questions is the possibility of persistent pattern formation/selforganization from a randomly
emitted distribution of brain waves influenced by inhomogeous anisotropic structure of their dispersion. Considering
a simplest case of fixed difference between wave excitation and dissipation, i.e. assuming propagation of wave packets
of fixed width (exponentially decaying) and searching for any closed parts of wave trajectories (loops) that are shorter
than the packet width can provide a clue about possible answer. Fig. 4 as well as Figs. A1 to A16 (and hyperlinked
movies) show formation of persistent loops for a variety of wave packet initial conditions as well as for different models
of conductivity tensor inhomogeneity and anisotropy constructed from dMRI or whole brain tractography.
Examples of wave trajectories and emergent persistent loop patterns for the spherical shell cortex model with
varying amounts of tensor anisotropy and inhomogeneous shell layer thickness (Figs. A1 to A4).
Examples of wave trajectories and emergent persistent loop patterns for cortical fold geometry with different
approaches used for estimation of inhomogeneity and anisotropy are shown in Figs.A5 to A16. Among those examples
are several simple cases with variable inhomogeneity and fixed anisotropy (similar to the above spherical shell cortex
model) as well as with more complex estimates of anisotropy based on multiple diffusion gradients MRI (dMRI)
acquisitions. Assuming linear dependence between diffusion and conductivity tensors [26] several combination of
the diffusion tensor eigenvectors were used to describe the conductivity tensor anisotropy.
In even more complex
approach, the full brain tractography [15] generated fiber distributions [25] were used to infer anisotropy through
direct integration of single fiber anisotropy parameters in each voxel.
All wave trajectory figures are hyperlinked and include references to location of movies showing dynamical devel-
opment of wave trajectories.
Complete wave trajectory
10
Emergent loop pattern
FIG. A1. An example of complete wave trajectory (top) and emergent loop pattern (bottom) for the spherical shell cortex model
with crossing fibers anisotropy tensor σ(3) and narrow inhomogeneity layer (n=0, r0=0.5 and r1=0.9). The trajectory was ini-
tialized with wave vector k = (−0.1/√2, 0.1/√2,−0.1) inside the inhomogeneous layer with voxel coordinates r=(142,142,142).
High resolution movie links: S1-H1/S1-H1, S1-H2/S1-H2. Low resolution movie links: S1-L1/S1-L1, S1-L2/S1-L2.
Complete wave trajectory
11
Emergent loop pattern
FIG. A2. An example of complete wave trajectory (top) and emergent loop pattern (bottom) for the spherical shell cortex
model with crossing fibers anisotropy tensor σ(3) and slightly wider inhomogeneity layer (n=2, r0=0.5 and r1=0.9). The
trajectory was initialized with wave vector k = (−0.1/√2, 0.1/√2,−0.1) inside the inhomogeneous layer with voxel coordinates
r=(80,33,100). High resolution movie links: S2-H1/S2-H1, S2-H2/S2-H2. Low resolution movie links: S2-L1/S2-L1, S2-L2/S2-
L2.
Complete wave trajectory
12
Emergent loop pattern
FIG. A3. An example of complete wave trajectory (top) and emergent loop pattern (bottom) for the spherical shell cortex
orientation single fiber anisotropy tensor σ(2) and wide inhomogeneity layer (n=25, r0=0.5 and r1=0.9). The
model with 45
trajectory was initialized with wave vector k = (−0.1/√2, 0.1/√2,−0.1) inside the inhomogeneous layer with voxel coordinates
◦
r=(145,145,145). High resolution movie links: S3-H1/S3-H1, S3-H2/S3-H2. Low resolution movie links: S3-L1/S3-L1, S3-
L2/S3-L2.
Complete wave trajectory
13
Emergent loop pattern
FIG. A4. An example of complete wave trajectory (top) and emergent loop pattern (bottom) for the spherical shell cortex
◦
orientation single fiber anisotropy tensor σ(2) and intermediately wide inhomogeneity layer (n=14, r0=0.5 and
model with 45
r1=0.9). The trajectory was initialized with wave vector k = (−0.1/√2, 0.1/√2,−0.1) inside the inhomogeneous layer with
voxel coordinates r=(141,141,141). High resolution movie links: S4-H1/S4-H1, S4-H2/S4-H2. Low resolution movie links:
S4-L1/S4-L1, S4-L2/S4-L2.
Complete wave trajectory
14
Emergent loop pattern
FIG. A5. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with diffusion MRI
derived position dependent anisotropy tensor RT σ(1)R (ε = (λ(2)
d ). The trajectory was initialized with wave
vector k=(-0.75, 0.23, -0.62) inside the inhomogeneous layer with voxel coordinates r=(55,130,130). High resolution movie
links: S5-H1/S5-H1, S5-H2/S5-H2, S5-H3/S5-H3. Low resolution movie links: S5-L1/S5-L1, S5-L2/S5-L2, S5-L3/S5-L3.
d )/2/λ(1)
d + λ(3)
Complete wave trajectory
15
Emergent loop pattern
FIG. A6. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with diffusion MRI
derived position dependent anisotropy tensor RT σ(1)R (ε = λ(2)
d ). The trajectory was initialized with wave vector k=(-
0.9,0.24,0.36) inside the inhomogeneous layer with voxel coordinates r=(66,94,126). High resolution movie links: S6-H1/S6-H1,
S6-H2/S6-H2, S6-H3/S6-H3. Low resolution movie links: S6-L1/S6-L1, S6-L2/S6-L2, S6-L3/S6-L3.
d /λ(1)
Complete wave trajectory
16
Emergent loop pattern
FIG. A7. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with diffusion MRI derived
position dependent anisotropy tensor RT σ(1)R (σ(1)
d ). The trajectory was initialized with
wave vector k=(-0.77,-0.63,0.11) inside the inhomogeneous layer with voxel coordinates r=(49,153,97). High resolution movie
links: S7-H1/S7-H1, S7-H2/S7-H2, S7-H3/S7-H3. Low resolution movie links: S7-L1/S7-L1, S7-L2/S7-L2, S7-L3/S7-L3.
11 = λ(3)
d /λ(1)
and σ(1)
22 = λ(2)
d /λ(1)
d
Complete wave trajectory
17
Emergent loop pattern
FIG. A8. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with fixed anisotropy
tensor σ(1) (ε = 0.1). The trajectory was initialized with wave vector k=(0.56,0.58,-0.59) inside the inhomogeneous layer with
voxel coordinates r=(72,156,120). High resolution movie links: S8-H1/S8-H1, S8-H2/S8-H2, S8-H3/S8-H3. Low resolution
movie links: S8-L1/S8-L1, S8-L2/S8-L2, S8-L3/S8-L3.
Complete wave trajectory
18
Emergent loop pattern
FIG. A9. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with diffusion MRI derived
position independent anisotropy tensor RT σ(1)R (ε = 0.1). The trajectory was initialized with wave vector k=(-0.13,-0.97,-
0.21) inside the inhomogeneous layer with voxel coordinates r=(111,103,110). High resolution movie links: S9-H1/S9-H1,
S9-H2/S9-H2, S9-H3/S9-H3. Low resolution movie links: S9-L1/S9-L1, S9-L2/S9-L2, S9-L3/S9-L3.
Complete wave trajectory
19
Emergent loop pattern
FIG. A10. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with fixed anisotropy
tensor σ(1) (ε = 0.01). The trajectory was initialized with wave vector k=(0.42,-0.41,0.81) inside the inhomogeneous layer
with voxel coordinates r=(114,85,22). High resolution movie links: S10-H1/S10-H1, S10-H2/S10-H2, S10-H3/S10-H3. Low
resolution movie links: S10-L1/S10-L1, S10-L2/S10-L2, S10-L3/S10-L3.
Complete wave trajectory
20
Emergent loop pattern
FIG. A11. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with fixed anisotropy
tensor σ(2) (ε = 0.1). The trajectory was initialized with wave vector k=(-0.90,0.24,0.36) inside the inhomogeneous layer
with voxel coordinates r=(66,86,126). High resolution movie links: S11-H1/S11-H1, S11-H2/S11-H2, S11-H3/S11-H3. Low
resolution movie links: S11-L1/S11-L1, S11-L2/S11-L2, S11-L3/S11-L3.
Complete wave trajectory
21
Emergent loop pattern
An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for
FIG. A12.
the cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with diffusion MRI
derived position independent anisotropy tensor RT σ(1)R (ε = 0.01). The trajectory was initialized with wave vector k=(-
0.90,0.24,0.36) inside the inhomogeneous layer with voxel coordinates r=(66,87,126). High resolution movie links: S12-H1/S12-
H1, S12-H2/S12-H2, S12-H3/S12-H3. Low resolution movie links: S12-L1/S12-L1, S12-L2/S12-L2, S12-L3/S12-L3.
Complete wave trajectory
22
Emergent loop pattern
anisotropy tensor 1/N(cid:80)
FIG. A13. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with tractography derived
k σ(1)Rk (ε = 0.001). The trajectory was initialized with wave vector k=(0.4,-0.83,0.4) inside
the inhomogeneous layer with voxel coordinates r=(75,61,90). High resolution movie links: S13-H1/S13-H1, S13-H2/S13-H2,
S13-H3/S13-H3. Low resolution movie links: S13-L1/S13-L1, S13-L2/S13-L2, S13-L3/S13-L3.
k RT
Complete wave trajectory
23
Emergent loop pattern
anisotropy tensor 1/N(cid:80)
FIG. A14. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with tractography derived
k σ(1)Rk (ε = 0.01). The trajectory was initialized with wave vector k=(0.66,0.5,0.56) inside
the inhomogeneous layer with voxel coordinates r=(33,77,56). High resolution movie links: S14-H1/S14-H1, S14-H2/S14-H2,
S14-H3/S14-H3. Low resolution movie links: S14-L1/S14-L1, S14-L2/S14-L2, S14-L3/S14-L3.
k RT
Complete wave trajectory
24
Emergent loop pattern
anisotropy tensor 1/N(cid:80)
FIG. A15. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with tractography derived
k σ(1)Rk (ε = 0.1). The trajectory was initialized with wave vector k=(0.36,0.67,-0.65) inside
the inhomogeneous layer with voxel coordinates r=(77,123,64). High resolution movie links: S15-H1/S15-H1, S15-H2/S15-H2,
S15-H3/S15-H3. Low resolution movie links: S15-L1/S15-L1, S15-L2/S15-L2, S15-L3/S15-L3.
k RT
Complete wave trajectory
25
Emergent loop pattern
anisotropy tensor 1/N(cid:80)
FIG. A16. An example of randomly initialized complete wave trajectory (top) and emergent loop pattern (bottom) for the
cortical fold model with inhomogeneity extracted from HRA volume registered to MNI152 space and with tractography derived
k σ(1)Rk (ε = 0). The trajectory was initialized with wave vector k=(-0.06,0.38,0.92) inside the
inhomogeneous layer with voxel coordinates r=(132,113,96). High resolution movie links: S16-H1/S16-H1, S16-H2/S16-H2,
S16-H3/S16-H3. Low resolution movie links: S16-L1/S16-L1, S16-L2/S16-L2, S16-L3/S16-L3.
k RT
26
[1] H. Monai, M. Inoue, H. Miyakawa, and T. Aonishi, Low-
frequency dielectric dispersion of brain tissue due to elec-
trically long neurites, Phys Rev E Stat Nonlin Soft Mat-
ter Phys 86, 061911 (2012); L. Ingber and P. L. Nunez,
Neocortical dynamics at multiple scales: EEG standing
waves, statistical mechanics, and physical analogs, Math
Biosci 229, 160 (2011).
[2] X. P. Li, Q. Xia, D. Qu, T. C. Wu, D. G. Yang, W. D.
Hao, X. Jiang, and X. M. Li, The Dynamic Dielec-
tric at a Brain Functional Site and an EM Wave Ap-
proach to Functional Brain Imaging, Scientific Reports
4, 6893 (2014); A. Zalesky, A. Fornito, L. Cocchi, L. L.
Gollo, and M. Breakspear, Time-resolved resting-state
brain networks, Proc. Natl. Acad. Sci. U.S.A. 111, 10341
(2014).
[3] D. C. Van Essen, S. M. Smith, D. M. Barch, T. E. J.
Behrens, E. Yacoub, K. Ugurbil, and for the WU-Minn
HCP Consortium, The WU-Minn Human Connectome
Project: An overview., NeuroImage 80, 62 (2013).
[4] D. S. Tuch, V. J. Wedeen, A. M. Dale, J. S. George,
and J. W. Belliveau, Conductivity Mapping of Biologi-
cal Tissue Using Diffusion MRI, Annals of the New York
Academy of Sciences 888, 314 (1999); J. Haueisen, D. S.
Tuch, C. Ramon, P. H. Schimpf, V. J. Wedeen, J. S.
George, and J. W. Belliveau, The influence of brain tis-
sue anisotropy on human EEG and MEG, NeuroImage
15, 159 (2002); H. Hallez, B. Vanrumste, P. V. Hese,
Y. D'Asseler, I. Lemahieu, and R. V. de Walle, A finite
difference method with reciprocity used to incorporate
anisotropy in electroencephalogram dipole source local-
ization, Physics in Medicine & Biology 50, 3787 (2005).
[5] P. L. Nunez and R. Srinivasan, Neocortical dynamics
due to axon propagation delays in cortico-cortical fibers:
EEG traveling and standing waves with implications for
top-down influences on local networks and white matter
disease, Brain Res. 1542, 138 (2014).
[6] L. Muller, G. Piantoni, D. Koller, S. S. Cash, E. Halgren,
and T. J. Sejnowski, Rotating waves during human sleep
spindles organize global patterns of activity that repeat
precisely through the night, Elife 5 (2016).
[7] E. V. Lubenov and A. G. Siapas, Hippocampal theta os-
cillations are travelling waves, Nature 459, 534 (2009);
H. Zhang, A. J. Watrous, A. Patel, and J. Jacobs, Theta
and Alpha Oscillations Are Traveling Waves in the Hu-
man Neocortex, Neuron 98, 1269 (2018); L. Muller,
F. Chavane, J. Reynolds, and T. J. Sejnowski, Cortical
travelling waves: mechanisms and computational princi-
ples, Nat. Rev. Neurosci. 19, 255 (2018).
[8] S. E. Fox, S. Wolfson, and J. B. Ranck, Hippocam-
pal theta rhythm and the firing of neurons in walking
and urethane anesthetized rats, Exp Brain Res 62, 495
(1986); M. Stewart, G. J. Quirk, M. Barry, and S. E.
Fox, Firing relations of medial entorhinal neurons to the
hippocampal theta rhythm in urethane anesthetized and
walking rats, ibid. 90, 21 (1992); A. Czurko, J. Huxter,
Y. Li, B. Hangya, and R. U. Muller, Theta phase classi-
fication of interneurons in the hippocampal formation of
freely moving rats, J. Neurosci. 31, 2938 (2011).
[9] S. A. Weiss and D. S. Faber, Field effects in the CNS
play functional roles, Front Neural Circuits 4, 15 (2010).
[10] C. Qiu, R. S. Shivacharan, M. Zhang, and D. M. Durand,
Can Neural Activity Propagate by Endogenous Electrical
Field?, J. Neurosci. 35, 15800 (2015); M. Zhang, T. P.
Ladas, C. Qiu, R. S. Shivacharan, L. E. Gonzalez-Reyes,
and D. M. Durand, Propagation of epileptiform activity
can be independent of synaptic transmission, gap junc-
tions, or diffusion and is consistent with electrical field
transmission, ibid. 34, 1409 (2014); C. C. Chiang, R. S.
Shivacharan, X. Wei, L. E. Gonzalez-Reyes, and D. M.
Durand, Slow periodic activity in the longitudinal hip-
pocampal slice can self-propagate non-synaptically by a
mechanism consistent with ephaptic coupling, J. Phys-
iol. (Lond.) 597, 249 (2019); R. S. Shivacharan, C. C.
Chiang, M. Zhang, L. E. Gonzalez-Reyes, and D. M. Du-
rand, Self-propagating, non-synaptic epileptiform activ-
ity propagates by endogenous electric fields, Exp. Neurol.
(2019).
[11] S. Gabriel, R. W. Lau, and C. Gabriel, The dielectric
properties of biological tissues: II. Measurements in the
frequency range 10 Hz to 20 GHz, Phys Med Biol 41,
2251 (1996); The dielectric properties of biological tis-
sues: III. Parametric models for the dielectric spectrum
of tissues, 41, 2271 (1996).
[12] J. W. S. Rayleigh, On Waves Propagated along the Plane
Surface of an Elastic Solid, Proceedings of the London
Mathematical Society 17, 4 (1885).
[13] I. N. Kartashov and M. V. Kuzelev, Dissipative surface
waves in plasma, Plasma Physics Reports 40, 650 (2014).
[14] V. L. Galinsky and L. R. Frank, Automated segmentation
and shape characterization of volumetric data, Neuroim-
age 92, 156 (2014).
[15] V. L. Galinsky and L. R. Frank, Simultaneous multi-scale
diffusion estimation and tractography guided by entropy
spectrum pathways, IEEE Trans. Med. Imag. 34, 1177
(2015).
[16] V. L. Galinsky and L. R. Frank, Symplectomorphic regis-
tration with phase space regularization by entropy spec-
trum pathways, Magn Reson Med , 1 (2018).
[17] A. Martinez, P. A. Gaspar, S. A. Hillyard, S. Bickel,
P. Lakatos, E. C. Dias, and D. C. Javitt, Neural oscil-
latory deficits in schizophrenia predict behavioral and
neurocognitive impairments, Front Hum Neurosci 9, 371
(2015); M. G. Woldorff, M. Liotti, M. Seabolt, L. Busse,
J. L. Lancaster, and P. T. Fox, The temporal dynamics
of the effects in occipital cortex of visual-spatial selective
attention, Brain Res Cogn Brain Res 15, 1 (2002).
[18] V. L. Galinsky and L. R. Frank, Brain Wave Loops
movies (2018), supplemental Material on FigShare and
DropBox.
[19] C. A. Anastassiou, R. Perin, H. Markram, and C. Koch,
Ephaptic coupling of cortical neurons, Nat. Neurosci. 14,
217 (2011).
[20] V. Fonov, A. C. Evans, K. Botteron, C. R. Almli, R. C.
McKinstry, D. L. Collins, W. S. Ball, A. W. Byars,
M. Schapiro, W. Bommer, A. Carr, A. German, S. Dunn,
M. J. Rivkin, D. Waber, R. Mulkern, S. Vajapeyam,
A. Chiverton, P. Davis, J. Koo, J. Marmor, C. Mrakot-
sky, R. Robertson, G. McAnulty, M. E. Brandt, J. M.
Fletcher, L. A. Kramer, G. Yang, C. McCormack,
K. M. Hebert, H. Volero, K. Botteron, R. C. McKinstry,
W. Warren, T. Nishino, C. R. Almli, R. Todd, J. Con-
stantino, J. T. McCracken, J. Levitt, J. Alger, J. O'Neil,
27
A. Toga, R. Asarnow, D. Fadale, L. Heinichen, C. Ire-
land, D. J. Wang, E. Moss, R. A. Zimmerman, B. Bintliff,
R. Bradford, J. Newman, A. C. Evans, R. Arnaoutelis,
G. B. Pike, D. L. Collins, G. Leonard, T. Paus, A. Zij-
denbos, S. Das, V. Fonov, L. Fu, J. Harlap, I. Leppert,
D. Milovan, D. Vins, T. Zeffiro, J. Van Meter, N. Lange,
M. P. Froimowitz, K. Botteron, C. R. Almli, C. Rainey,
S. Henderson, T. Nishino, W. Warren, J. L. Edwards,
D. Dubois, K. Smith, T. Singer, A. A. Wilber, C. Pier-
paoli, P. J. Basser, L. C. Chang, C. G. Koay, L. Walker,
L. Freund, J. Rumsey, L. Baskir, L. Stanford, K. Sirocco,
K. Gwinn-Hardy, G. Spinella, J. T. McCracken, J. R.
Alger, J. Levitt, and J. O'Neill, Unbiased average age-
appropriate atlases for pediatric studies, Neuroimage 54,
313 (2011).
[21] V. Fonov, A. Evans, R. McKinstry, C. Almli, and
D. Collins, Unbiased nonlinear average age-appropriate
brain templates from birth to adulthood, NeuroImage
47, S102 (2009), organization for Human Brain Mapping
2009 Annual Meeting.
[22] M. Niethammer, R. San Jose Estepar, S. Bouix, M. Shen-
ton, and C. F. Westin, On diffusion tensor estimation,
Conf Proc IEEE Eng Med Biol Soc 1, 2622 (2006).
[23] J. M. Gomes, C. Bedard, S. Valtcheva, M. Nelson,
V. Khokhlova, P. Pouget, L. Venance, T. Bal, and
A. Destexhe, Intracellular Impedance Measurements Re-
veal Non-ohmic Properties of the Extracellular Medium
around Neurons, Biophys. J. 110, 234 (2016).
[24] B. Barbour, Analysis of Claims that the Brain Extracel-
lular Impedance Is High and Non-resistive, Biophys. J.
113, 1636 (2017).
[25] V. L. Galinsky and L. R. Frank, The lamellar structure
of the brain fiber pathways, Neural Comput. 28, 2533
(2016).
[26] D. S. Tuch, V. J. Wedeen, A. M. Dale, J. S. George, and
J. W. Belliveau, Conductivity tensor mapping of the hu-
man brain using diffusion tensor MRI, Proc. Natl. Acad.
Sci. U.S.A. 98, 11697 (2001).
|
1606.05428 | 1 | 1606 | 2016-06-17T06:57:43 | How does a protein reach its binding locus: sliding along DNA chain or not? | [
"physics.bio-ph",
"q-bio.SC"
] | In gene expression, various kinds of proteins need to bind to specific locus of DNA. It is still not clear how these proteins find their target locus. In this study, the mean first-passage time (FPT) of protein binding to its target locus on DNA chain is discussed by a chain-space coupled model. Our results show that the 1-dimensional diffusion constant has a critical value, with which the mean time spent by a protein to find its target locus is almost independent of the binding rate of protein to DNA chain and the detachment rate from DNA chain. Which implies that, the frequency of protein binding to DNA and the sliding time on DNA chain have little influence on the search efficiency, and therefore whether or not the 1-dimensional sliding on DNA chain increases the search efficiency depends on the 1-dimensional diffusion constant of the protein on DNA chain. This study also finds that only protein bindings to DNA loci which are close to the target locus help to increase the search efficiency, while bindings to those loci which are far from the target locus might delay the target binding process. As expected, the mean FPT increases with the distance between the initial position of protein in cell space and its target locus on DNA chain. The direct binding probability, which can be regarded as one index to describe if the 1-dimensional sliding along DNA chain is helpful to increase the search efficiency is calculated. Our results show that the influence of 1-dimensional sliding along DNA chain on the search process depends on both diffusion constants of protein in cell space and on the 1-dimensional DNA chain. | physics.bio-ph | physics | How does a protein reach its binding locus: sliding along DNA chain or not?
Shanghai Key Laboratory for Contemporary Applied Mathematics, Centre for Computational Systems Biology,
Jingwei Li, Yunxin Zhang
School of Mathematical Sciences, Fudan University, Shanghai 200433, China.
(Dated: September 3, 2018)
In gene expression, various kinds of proteins (such as polymerase or transcription factor) need
to bind to specific locus of DNA. Although sophisticated experiments have been done according to
this process, it is still not clear how these proteins find their target locus. Are these target-search
processes completed mainly by 3-dimensional diffusion in cell space or with the aid of 1-dimensional
sliding along DNA chain? Previous studies have shown that sliding along DNA chain may help to
increase the search efficiency. While recent experiments also found that the length of DNA sequence
has little influence on the search time. In this study, the mean first-passage time (FPT) of protein
binding to its target locus on DNA chain is discussed by a chain-space coupled model. In which
the cell space is simply represented by a 2-dimensional rectangular lattice and the DNA chain is
simplified to a 1-dimensional lattice with length L. Our results show that the mean FPT has power
law relation with the 2-dimensional diffusion constant approximately. The 1-dimensional diffusion
constant has a critical value, with which the mean time spent by a protein to find its target locus is
almost independent of the binding rate of protein to DNA chain and the detachment rate from DNA
chain. Which implies that, the frequency of protein binding to DNA and the sliding time on DNA
chain have little influence on the search efficiency, and therefore whether or not the 1-dimensional
sliding on DNA chain increases the search efficiency depends on the 1-dimensional diffusion constant
of the protein on DNA chain. This study also finds that only protein bindings to DNA loci which
are close to the target locus help to increase the search efficiency, while bindings to those loci which
are far from the target locus might delay the target binding process. As expected, the mean FPT
increases with the distance between the initial position of protein in cell space and its target locus
on DNA chain. While our results show that the mean FPT does not change monotonically with
the distance between the initial position of protein and the DNA chain. To know how a protein
reaches its target locus, i.e., binding the target through its adjacent loci of DNA or directly binding
through its nearest neighbor position in the cell space, the direct binding probability, which can be
regarded as one index to describe if the 1-dimensional sliding along DNA chain is helpful to increase
the search efficiency is calculated. Our results show that the influence of 1-dimensional sliding along
DNA chain on the search process depends on both diffusion constants of protein in cell space and
on the 1-dimensional DNA chain.
Keywords: gene expression; first-passage time, first-passage probability, RNA polymerase.
6
1
0
2
n
u
J
7
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
8
2
4
5
0
.
6
0
6
1
:
v
i
X
r
a
I.
INTRODUCTION
During gene expression, specific protein molecules, such as RNA polymerase and transcription factor, need to
recognize and bind to certain loci on DNA chain, which usually lie in promoter domain [1 -- 7]. These binding processes
are important for biological systems to regulate gene expressions [8 -- 12]. The mechanism of how a protein reaches
2
FIG. 1: Diagram of the chain-space coupled model used in this study. The DNA chain is simplified as a one-dimensional lattice
with length L (L = 5 in this diagram), and assumed to lie at the middle of the cell space. The cell space is simplified as a
rectangular two-dimensional lattice with size M × N (M = 7, N = 5 here). The target locus of a protein is assumed to lie at
the center of the DNA chain, i.e. the lattice site (L + 1)/2 for an odd number L. A protein molecule in cell space can walk
randomly between adjacent lattice sites with rate D2, or along the DNA chain with rate D1. Proteins can bind to or detach
from the DNA chain randomly with rate kb and kd, respectively. The binding rate of a protein molecule from cell space to the
target locus on DNA chain is denoted by kt.
its target locus on DNA is a basic biophysical problem, and has been extensively studied both experimentally and
theoretically [13 -- 16]. Nevertheless, the mechanism of this target search process remains unclear [16, 17]. In references,
various methods of theoretical analysis have been presented to try to explain this fast search process in cells [18 --
20], which is usually called facilitated diffusion (FD) due to its high efficiency. Including the approach of lowering
dimensionality [15, 17, 21, 22], electrostatic effects [23], correlations between 3D and 1D motions [16, 24, 25], transitions
between different chemical states [14, 26], as well as bending fluctuations and hydrodynamics [27].
In this study, we mainly want to show that if the one-dimensional sliding of protein along DNA chain attributes to
this search process for a target on DNA. Or in other words, if the one-dimensional sliding is essential to increase the
search efficiency, and can shorten the search time effectively. In previous studies [17, 21, 28 -- 38], the protein molecule
is thought to first bind to a nonspecific locus of DNA chain through three-dimensional diffusion in cell space, and
then slide along the one-dimensional DNA chain to reach the target locus (or binding site). Corresponding theoretical
analysis showed that, with the help of this one-dimensional sliding along DNA chain (lowering dimensionality), the
target search efficiency of a protein can be greatly increased [15, 17, 21, 22]. So it is believed that the one-dimensional
sliding along DNA chain is essential to accelerate this search process, and consequently is important to gene expression
in cell. However, observations in recent experiments [23, 39, 40] showed that most proteins reached their target loci
without long-range one-dimensional sliding along the DNA chain, and therefore this target search process is mainly
completed through three-dimensional diffusion in cell space. So, the mechanism of how proteins in cells, including
polymerase and transcription factor, can find their binding sites on DNA chain effectively remains unclear.
One can image that the search mode of proteins depends on both cell environment and protein properties, especially
the diffusion constants in cell space and along DNA chain, as well as the binding/unbinding rates of proteins to/from
the DNA chain. For example, with low values of one-dimensional diffusion constant along DNA chain and high values
of unbinding rate, proteins will reach their target loci mainly through diffusion in cell space, and vice versa. Thus
the contribution of one-dimensional sliding along DNA chain to this target search process of proteins is influenced by
both the cell environment and protein properties.
3
To show how the search mode of a protein molecule is influenced by diffusion constants and binding/unbinding
rates, a similar model as the one used in [41] is employed in this study. In which the DNA chain is simplified to a
one-dimensional lattice with length L, with each DNA locus represented by a lattice site. The cell space is simplified
to a two-dimensional rectangular lattice with M × N lattice sites, see Fig. 1. Five parameters are included in our
model, one-dimensional diffusion constant along DNA chain D1, two-dimensional diffusion constant in cell space D2,
binding/unbinding rate of protein molecule to/from nonspecific binding site kb and kd, and direct binding rate from
cell space to the target locus on DNA chain kt. If the diffusion constant D2 is extremely large compared with binding
rates kt and kb, the mean first-passage time (FPT) of a protein molecule from cell space to target binding site can be
obtained explicitly. For general cases, numerical computations are employed.
Our results show that there is a critical value of the diffusion constant D1, with which the mean FPT is insensitive
to values of binding/unbinding rates kb and kd, unless kb and 1/kd are extremely large. In other words, with this
critical value, how often and how long a protein molecule slides along the DNA chain have almost no influence on the
target search efficiency. For values of diffusion constant D1 which is larger than this critical value, the one-dimensional
sliding along DNA chain will be helpful to increase the search efficiency, and vice versa. Which means that for large
values of D1, the search efficiency will increase with the binding rate kb while decrease with the unbinding rate kd.
While for small values of D1, search efficiency will decrease with rate kb but increase with rate kd. Meanwhile, in this
study, the influences of two-dimensional diffusion constant D2, the length L of DNA chain, and the distance dtarget
(dchain) between the initial position of a protein molecule and its target locus (DNA chain) are also discussed. Finally,
a method to calculate the probability P direct that a protein molecule reaches the target DNA locus directly, i.e. bind
to the target locus through the adjacent positions in cell space but not from its nearest neighboring loci on DNA
chain, is also presented. Here, P direct can be regarded as one index to evaluate the importance of the one-dimensional
sliding along DNA chain in the target search process of proteins.
This paper is organized as follows. The model used in this study will be briefly introduced in Section II, and
then theoretical methods to get the mean FPT for large limit values of diffusion constant D2 will be presented in
Section III. For general cases, results obtained by numerical computations will be given in Section IV. In Section V,
the dependence of direct binding probability P direct on diffusion constants D1 and D2 will be discussed theoretically.
Finally, concluding and remarks will be presented in the last section.
II. MODEL DESCRIPTION
In our model, the DNA chain is simplified to a one-dimensional lattice with length L, with each DNA locus
represented by a lattice site. While the cell space is simplified to a two-dimensional rectangular lattice with size
M × N , see Fig. 1 for a schematic depiction in which L = 5 and M = 7, N = 5. The diffusion of protein molecule is
then simplified to random walks between adjacent lattice sites. For convenience, this study assumes that the target
DNA locus lies at the center of the DNA chain, i.e. the lattice site (L + 1)/2 of DNA chain. Meanwhile, the DNA
chain is assumed to be horizontal and lies at the center of the rectangular lattice (in this study, L, M , N are always
chosen to be odd numbers).
To reach its target binding site on DNA chain, the protein molecule first walks randomly on the two-dimensional
rectangular lattice with rate D2. When it reaches lattice sites adjacent to the DNA chain, it may either bind to the
4
nearest DNA site with rate kb if this site is not the target locus, or bind with rate kt if the nearest site is the target
locus, or just walks away randomly with rate D2 in the rectangular lattice. The observations in recent experiments
[39] showed that the binding rate of RNA polymerase to promotor is usually larger than that to other regions of DNA
chain, therefore kt is generally larger than kb. After binding to DNA chain, the protein molecule will walk randomly
along the one-dimensional lattice with jumping rate D1. At sites which are not the target DNA locus, the protein
may detach into the cell space (i.e. the rectangular lattice) again with rate kd.
In this study, periodic boundary conditions are used for random walk in the two-dimensional rectangular lattice,
which means that the left and bottom boundaries of the rectangular lattice are connected with the right and top
boundaries, respectively.
In all the following numerical calculations, M = N = 101 are used, and each site of
the rectangular lattice is denoted by its position coordinate (i, j), with i the column index and j the row index of
rectangular lattice, see Fig. 1.
III. MEAN FIRST-PASSAGE TIME FOR LARGE LIMIT VALUES OF RATE D2
In this section, we will derive the expression of mean FPT for the special cases in which the two-dimensional
diffusion rate D2 is large enough. For convenience, we define several notations as follows. Let Ui be the mean FPT
of a protein which initiates from site i of DNA chain to reach the target site (L + 1)/2. Let Qi be the splitting
probability that a protein initiated at DNA site i reaches the target site without unbinding from the DNA chain, and
Ti be the conditional mean FPT of Qi. Let Pij be the splitting probability that a protein initiated at DNA site i
detaches from the DNA site j without reaching the target site. It is obvious that Qi + Pj6=(L+1)/2 Pij = 1. Let Si
be the conditional mean FPT of 1 − Qi. Actually, Si is the conditional mean FPT that a protein initiated at DNA
site i detaches from the chain (at any site) without reaching the target site. Let Ri be the mean FPT of a protein
from position ((M − L)/2 + i, (N + 1)/2) (corresponding to the position of DNA site i) in the 2D rectangular lattice
to DNA chain. Let Oij be the splitting probability that a protein initiated at position ((M − L)/2 + i, (N + 1)/2)
binds to the DNA chain at site j. For definitions of (conditional) mean FPT, splitting probability, see [42].
With the above notations, one can show that
Ui = QiTi + (
L
X
j=1
Pij )Si +
L
X
j=1
PijRj +
L
X
k=1
L
X
j=1
Pij OjkUk,
where 1 ≤ i ≤ L. Or it can be written as the following matrix form,
U = Q ◦ T + (P e) ◦ S + P R + P OU.
(1)
(2)
Here, U , Q, T , S, R are L × 1 vectors with components Ui, Qi, Ti, Si, Ri respectively. P and O are L × L matrices
with components Pij and Oij . All components of the L × 1 vector e equal one. Q ◦ T means an L × 1 vector with
components QiTi. (P e) ◦ S is defined in the same way.
Let Fi be the mean FPT of a protein from position ((M − L)/2 + i, (N + 1)/2) in cell space to the target site
(L + 1)/2 of DNA chain. Then we have
Fi = Ri +
L
X
j=1
Oij Uj.
(3)
5
Or in matrix form F = R + OU , with F an L × 1 vector with components Fi.
Given Q, T , P , S, R and O, the mean FPT U can be obtained from Eq. (1). But explicit expressions of O and
R are difficult to obtain. Meanwhile, even if O and R are obtained, one still need to solve the inverse of an L × L
matrix to get U . Therefore, in this section we only discuss the limit cases in which the two-dimensional diffusion rate
D2 is much larger than binding rates kb and kt of protein to DNA chain, i.e. D2 ≫ kb, kt. General cases will be
discussed in the next section by numerical computations. For these limit cases, one can easily show that R ≈ eR0
with R0 = (¯kbL/M N )−1 = CM N /(Lkb), Oij = C/L for j 6= (L + 1)/2, and Oij = Ckt/(Lkb) for j = (L + 1)/2.
Where ¯kb = [(L − 1)kb + kt]/L is the average binding rate of a protein to DNA chain, and C = kb/¯kb. Note that
U(L+1)/2 = 0, thereby
OU = (C/L)EU = C ¯U e,
(4)
where E is an L × L matrix with all elements equal to one, and ¯U = (PL
show that P e = e − Q, so
i=1 Ui)/L. From definitions one can easily
U = Q ◦ T + (P e) ◦ S + (e − Q)R0 + C(e − Q) ¯U ,
and
It can be proved that (see Section ?? of the Supplementary Material)
F = (R0 + C ¯U )e.
Thus,
Q ◦ T + (P e) ◦ S =
1
h
(e − Q).
U =
1
h
(e − Q) + (e − Q)R0 + C(e − Q) ¯U .
(5)
(6)
(7)
(8)
Define H := eT (e − Q)/L = e − Q, where eT is the transpose of e, i.e. is a 1 × L vector with all components equal
to one.
In fact, H is the average value of all components of vector e − Q, and therefore also denoted by e − Q.
Multiplying both sides of Eq. (8) by eT /L, we obtain
so
Substituting Eq. (10) into Eq. (6), we get
¯U =
H
h
+ R0H + CH ¯U ,
¯U =
H
h + R0H
1 − CH
.
F = (
R0 + C
h H
1 − CH
)e.
(9)
(10)
(11)
For the calculations of Q and H, see Section ?? of the Supplementary Material.
Let Ti,j be the mean FPT of a protein from position (i, j) in cell space to the target binding site (L + 1)/2 on DNA
chain. Define
¯T = PM
i=1 PN
M N
j=1 Ti,j
,
(12)
6
T1,1
T(M +1)/2,1
T1,(N +1)/2
T(M +1)/2,(N +1)/2
¯T
Large D2 limit
104
103
T
P
F
n
a
e
M
102
100
101
102
103
104
105
D2
FIG. 2: The average of mean FPT ¯T (see Eq. (12) for the definition of ¯T ), and four typical examples of mean FPT Ti,j ,
with (i, j) = (1, 1), ((M + 1)/2, 1), (1, (N + 1)/2), ((M + 1)/2, (N + 1)/2), respectively, as functions of two-dimensional diffusion
constant D2. In calculations, M = N = 101, L = 51, kb = 1, kd = 1, kt = 10, and D1 = 1 are used. One can find that all
numerical values of ¯T and Ti,j tend to the theoretical one as D2 → ∞, see discussions in Section III.
as the average of mean FPTs over the 2D space with prior uniform distribution.
In order to validate Eq. (11),
examples of T1,1, T(M+1)/2,1, T1,(N +1)/2, T(M+1)/2,(N +1)/2, and ¯T are numerically calculated and plotted in Fig. 2.
The results show that they all tend to the theoretical value given by Eq. (11) as rate D2 → ∞.
IV. MEAN FIRST-PASSAGE TIME TO REACH TARGET BINDING SITE: GENERAL CASES
In previous section, the mean FPT of a protein molecule to find its target binding site in DNA chain has been
obtained explicitly for large value limit of diffusion rate D2. In the following, we will discuss the general cases but
through numerical computations, see Section ?? of the Supplementary Material for some details of the numerical
method used in this study. We will mainly focus on the influences of diffusion rates D1 and D2, as well as the length
L of DNA chain. Meanwhile, the dependences of mean FPT on target distance dtarget and chain distance dchain
are also obtained numerically. Here dtarget/chain is the distance between initial position of protein molecule and the
target binding site/DNA chain. In the following subsection, we will first show that there exists a critical value of the
one-dimensional diffusion rate D1, with which the mean FPT is almost independent of the binding rate kb and the
unbinding rate kd.
A. The critical value of one-dimensional diffusion rate D1
The average of mean FPTs ¯T as functions of D1, kb and 1/kd are plotted in Fig. 3. Fig. 3(a) shows that, for
small values of diffusion rate D1, ¯T increases with binding rate kb. While for large values of D1, ¯T decreases with kb.
Which means that one-dimensional sliding along DNA chain helps to increase the target-search efficiency of a protein
only when its sliding speed on DNA chain is large enough. The plots in Fig. 3(b) show that, ¯T always decreases with
7
9500
¯T
9000
(a)
8500
0
9500
¯T
9000
(c)
8500
0
9500
¯T
9000
2
4
kb
6
8
(b)
10
8500
0
0.005
0.01
0.015
D1
0.02
0.025
0.03
9500
¯T
9000
2
4
6
8
1/kd
(d)
10
8500
0
0.005
0.01
0.015
D1
0.02
0.025
0.03
FIG. 3: The average of mean FPT ¯T of a protein in cell space to find its target locus on DNA chain, as functions of binding
rate kb (a), diffusion rate D1 on DNA chain (b,d), and the inverse of detachment rate, 1/kd (c).
In (a,c), the values of
D1 corresponding to curves from the top down are 0 to 0.03 with increment 0.001. In (b) the values of kb corresponding to
curves from the bottom up (according to the order at D1 = 0) are 0 to 10 with increment 0.1. While in (d) the values of
1/kd corresponding to curves from the bottom up (according to the order at D1 = 0) are 0 to 10 with increment 0.1. Other
parameter values used in calculations are M = N = 101, L = 51, D2 = 1, kt = 10, and kd = 1 in (a,b), kb = 1 in (c,d). The
plots in (b,d) show that there exists a critical value of D1, at which the average of mean FPT ¯T is insensitive to binding rate
kb and detachment rate kd.
the diffusion rate D1. But there exists a critical value D∗
1 the binding of protein to DNA chain will
decrease the search efficiency, and vice versa. Similar results can be obtained from the plots in Fig. 3(c,d). Which
show that for small values of D1, ¯T decreases with unbinding rate kd, while for large values of D1, ¯T increase with
kd. Therefore, all the plots in Fig. 3 imply that whether or not one-dimensional sliding along DNA chain is helpful
1, for D1 < D∗
to the target-search process depends on the values of diffusion rate D1. It is helpful only when D1 is larger than the
1, the average of mean FPT ¯T is almost insensitive to the change of binding rate kb
critical value D∗
and unbinding rate kd, see also plots in Fig. 5(c). In other words, when D1 takes this critical value, how often and
1. For D1 = D∗
how long a protein slides along the DNA chain have little influence on the time spent by it to reach its target binding
site on DNA. Note that we keep the two-dimensional diffusion rate D2 = 1 in all plots of Fig. 3. Obviously the value
of D∗
1 will increase with D2.
To know if the properties obtained for the average of mean FPT ¯T from Fig. 3 hold for mean FPT Ti,j, similar
figures for typical mean FPT Ti,j are plotted in Figs. ??-??, and Figs. ??(c)-??(c), with (i, j) = (1, 1), ((M +
1)/2, 1), (1, (N + 1)/2), and ((M + 1)/2, (N + 1)/2) respectively. From these plots we can conclude that Ti,j, the
mean FPT of a protein initiated at position (i, j) in the cell space to find its target site on DNA chain, has the same
properties as those of ¯T . The plots in Figs. ??-?? show that, for mean FPT T(M+1)/2,(N +1)/2, whose initial position
is exactly the same as the target site on DNA chain (but lies in cell space, see Fig. 1), the corresponding critical value
1 = 0.00752. While for mean FPTs T1,1, T(M+1)/2,1, T1,(N +1)/2, the critical value is D∗
of D1 is D∗
1 = 0.0151, which is
the same as that obtained from the average of mean FPT ¯T . This implies a fact that the properties of ¯T are mainly
8
x 104
(a)
6
5
4
3
¯T
2
1
0
100
x 104
5
(c)
4
3
¯T
2
1
0
100
x 1011
(b)
4
3
2
¯T
1
105
kb
1010
0
0
x 1011
4
(d)
2
4
kb
6
8
10
x 109
3
2
¯T
1
0
0
105
1/kd
1010
2
4
6
8
1/kd
10
x 109
FIG. 4: Typical examples of ¯T as functions of kb (a,b) and 1/kd (c,d), 1(cid:13) monotonically increase, 2(cid:13) decrease followed by
increase, and 3(cid:13) monotonically decrease. In (a,c), values of D1 used in calculations are D1 = 0.005, 0.03, 10 (from the top
down) respectively. (b,d) are limit cases with D1 = 0. Other parameter values used in calculations are M = N = 101, L = 51,
D2 = 1, kt = 10, and kd = 1 in (a,b), kb = 1 in (c,d). (a,c) show that ¯T tends to a constant dependent on D1 as kb → ∞ or
kd → 0.
determined by the mean FPTs Ti,j with initial positions (i, j) near the boundary of the cell space. This fact is due to
the effect of high dimension that for a given bounded domain, most of its points are closer to the domain boundary
than to the domain center [43]. This high dimension effect appears in other properties of mean FPT as well.
Finally, we want to point out that the above results about the critical value of D1 is valid approximately only when
the binding rate kb is not too large and unbinding rate kd is not too small. In fact, the plots in Fig. 4(a) show that,
if D1 is small/large, then ¯T will monotonically increase/decrease with kb, while for intermediate values of D1, ¯T first
decreases then increases with kb. Nevertheless, as long as D1 6= 0, ¯T will always tend to a constant as kb → ∞. For
the special cases with D1 = 0, ¯T increases linearly with kb, see Fig. 4(b). For the unbinding rate kd, similar results
can be obtained, see Fig. 4(c,d).
B. Behaviors of mean FPT within wide range of diffusion rates
In calculations of previous subsection, to show the existence of critical value of one-dimensional diffusion rate D1,
we always fixed the value of two-dimensional diffusion rate D2 = 1, and varied the one-dimensional diffusion rate D1
in an appropriate range. In this subsection, we will show how the average of mean FPT ¯T changes with D1 and D2
within a large range, i.e., with change from a small value to an extremely large value.
In Fig. 5(a), we give a logarithm-logarithm plot of the average mean FPT ¯T as a function of D2 with different
values of D1. The value of D2 changes in interval [10−5, 105]. For the two limit cases, i.e., with D1 = 0 and D1 = 1010
(here we use D1 = 1010 to show the large D1 limit properties), the corresponding curves, denoted by ¯TD1=0 and
¯TD1=∞ for convenience, parallel with each other in the logarithm-logarithm scale. Which means that ¯TD1=0/ ¯TD1=∞
is independent of diffusion rate D2. Actually, ¯TD1=0 can be regarded as the average of mean FPT Ti,j of a protein
9
9100
9050
¯T
9000
(b)
105
8950
1
11 21 31 41 51 61 71 81 91 101
L
100
D2
1
0.8
0.6
t
c
e
r
i
d
P
0.4
0.2
0
10−5
(d)
100
D2
105
2
4
6
8
10
1/kd
1010
¯T
105
(a)
100
10−5
9200
9100
¯T
9000
8900
8800
0
(c)
FIG. 5: (a) The value of ¯T as a function of D2 with different values of D1. From the top down, the values of D1 used in
calculations are D1 = 0, 10−4, 10−3, 10−2, 10−1, 1, 10, 102, 103, 1010 respectively. Other parameter values used in calculations
are M = N = 101, L = 51, kb = 1, kd = 1, and kt = 10. (b) Typical examples of ¯T as functions of length L of DNA chain.
Parameter values used in calculations are M = N = 101, D2 = 1, kb = 1, kd = 1, kt = 10, and D1 = 0, 0.005, 0.01 (from the
top down) respectively. For very small D1 (for example D1 = 0 in the figure), ¯T increases monotonically with L. While for
relatively large D1, ¯T decreases rapidly with small values of L, and then increases gradually with large L. (c) ¯T as functions
of 1/kd, with M = N = 101, L = 51, D2 = 1, kt = 10, D1 = 0.0151. The values of kb corresponding to curves from the
bottom up (according to the order near 1/kd = 2) are 0 to 10 with increment 1. These plots show that ¯T is insensitive to
kb and 1/kd in [0, 10] when D1 = 0.0151, see also Figs. 3(b,d). (d) Examples of probability P direct that a protein reaches
its target locus on DNA chain through direct binding from cell space but not from the adjacent binding sites of DNA, as
functions of two-dimensional diffusion rate D2, with parameter values M = N = 101, L = 51, kb = 1, kd = 1, kt = 10, and
D1 = 10−2, 10−1, 1, 10, 100 (from the top down) respectively. These plots show that P direct increases with D2 while decreases
with D1.
to reach a target site in cell space, while ¯TD1=∞ can be regarded as the average of Ti,j of a protein to reach a DNA
chain with L binding sites.
The plots in Fig. 5(a) also show that ¯T decreases with two-dimensional diffusion rate D2, and tends to constant
as D2 → ∞ while tends to infinity as D2 → 0. As D1 → 0, ¯T approaches ¯TD1=0 from below, but first for large values
of D2 and then for small values of D2. As D1 → ∞, ¯T approaches ¯TD1=∞ from above, while first for small values of
D2 and then for large values D2.
As functions of rates D1 and D2, mean FPTs T1,1, T(M+1)/2,1, T1,(N +1)/2 have similar properties as those of
¯T , see Fig. ??(a), Fig. ??(a), and Fig. ??(a). However, the plots in Fig. ??(a) show different behavior of
mean FPT T(M+1)/2,(N +1)/2. For D1 = 0, T(M+1)/2,(N +1)/2 is a constant, and is independent of D2. This can be
proved theoretically, see Section ?? of the Supplementary Material. For D1 = 1010, T(M+1)/2,(N +1)/2 tends to a
constant both as D2 → +∞ and D2 → 0. For large D1 limit, the protein bound to DNA chain will find its target
binding site instantaneously. Therefore, for these cases, T(M+1)/2,(N +1)/2 can be regarded as the FPT of a protein
at position ((M + 1)/2, (N + 1)/2) to find the DNA chain. So different from the limit case D1 = 0, for large D1
limit T(M+1)/2,(N +1)/2 increases with the two-dimensional diffusion rate D2. For D2 → 0, the protein at position
((M + 1)/2, (N + 1)/2) will reach the DNA chain mainly by binding to the target binding site. While for D2 → ∞,
the protein will reach each of the site of DNA chain equally. The curves of T(M+1)/2,(N +1)/2 for intermediate values
of D1 approach those of D1 = 0 and D1 = 1010 in a same way as previous discussed for ¯T .
In above discussions, we avoid the special cases with D2 = 0 since it is a singular point such that T(M+1)/2,(N +1)/2 =
1/kt, while the mean FPTs with other initial positions are infinity.
10
C. The influence of DNA chain length L
The average of mean FPTs ¯T as a function of length L of DNA chain is plotted in Fig. 5(b), with D1 = 0, 0.005, 0.01
respectively. There are two cases. For very small values of D1 (see the case with D1 = 0 in Fig. 5(b), we claim that
this is not the only case but just a typical one with very small D1 value), ¯T increases with L monotonically. Which
implies that binding to DNA chain always hinders the search of the target site. On the other hand, for relatively
large values of D1 (see the cases with D1 = 0.005, 0.01 in Fig. 5(b)), ¯T decreases rapidly for small values of L,
and then increases gradually for large values of L. An intuitionistic explanation is that, for small values of L, the
increase of chain length gives the protein more chances to bind to DNA chain near the target binding site, thereby
increases the search efficiency. For large values of L, although bindings to DNA chain happen more frequently, most
of them are far from the target site, thereby will decrease the search efficiency. One can image that the optimal length
L∗ of DNA chain, with which the average ¯T of mean FPTs reaches its minimum, increases with one-dimensional
diffusion constant D1. It has been experimentally found that, by shortening the flanking DNA (the part without
target promoter binding site), the rate of promoter binding does not change significantly [39]. This may due to the
negative but slight influence of protein bindings to DNA sites far from the target promoter.
Similar results can be obtained for the typical examples of mean FPTs T1,1, T(M+1)/2,1, T1,(N +1)/2, and
T(M+1)/2,(N +1)/2, see Figs. ??(b), ??(b), ??(b), and ??(b) respectively.
D. Mean FPT as functions of distances dtarget and dchain
Examples of scatter diagram of the mean FPT as a function of dtarget with different values of D1 and D2 are plotted
in Fig. 6. Roughly speaking, the mean FPT T increases with the distant dtarget. Except the special cases where D1 is
large while D2 is small, see Fig. 6(g). Since for these cases, the sliding of protein on DNA chain is fast, and therefore
the mean FPT is considerably influenced by the binding process of protein to DNA chain. From the plots in Fig. 6,
one can find that although the value of mean FPT T changes with diffusion rates D1 and D2, the shape of function
T (dtarget) does not change significantly.
One can image that if the one-dimensional sliding on DNA chain can decrease the search time remarkably, then
the mean FPT T of a protein will strongly depend on the distance dchain between the initial position of pro-
tein and the DNA chain. Here dchain is defined as the minimum of the distances between the initial position
of protein and all binding sites of the DNA chain. The plots in Fig. 7 show that the mean FPT T does not
increase monotonically with the distance dchain. However, one can find that the maximal value Tmax and min-
imal value Tmin, as well as the average value Taverage of mean FPT T do increase with distance dchain. Where
Tmax(dchain) := max{Ti,jthe distance between position (i, j) and DNA chain is dchain}, and Tmin and Taverage are de-
11
70.7107
70.7107
70.7107
70.7107
0
0
70.7107
1500
1000
(h)
70.7107
500
0
dtarget
(i)
70.7107
650
0
FIG. 6: The mean FPT Ti,j as a function of distance dtarget between initial position (i, j) and target locus on DNA. The
diffusion constant D1 used in calculations of each row of subfigures is D1 = 0.1, 1, 10 respectively (from the top down), and D2
used in each column is D2 = 1, 10, 100 (from left to right). Except the special cases with high values of D1 and low values of
D2, see subfigure (g), the mean FPT increases with dtarget roughly, and tends to a constant for large values of distance dtarget.
10000
8000
6000
4000
2000
0
0
10000
T
P
F
n
a
e
M
8000
6000
4000
2000
0
0
5000
2500
(a)
(d)
(g)
0
0
10000
8000
6000
4000
2000
0
0
10000
T
P
F
n
a
e
M
8000
6000
4000
2000
0
0
5000
2500
(a)
(d)
(g)
0
0
2000
1800
1600
1400
1200
70.7107
1000
0
2000
1500
1000
500
(b)
(e)
2000
1800
1600
1400
1200
55.9017
1000
0
2000
1500
1000
500
(b)
(e)
1100
1080
1060
1040
1020
70.7107
1000
0
1000
(c)
(f)
980
960
940
920
900
0
750
700
1100
1080
1060
1040
1020
55.9017
1000
0
1000
(c)
(f)
980
960
940
920
900
0
750
700
55.9017
55.9017
55.9017
55.9017
0
0
55.9017
1500
1000
(h)
55.9017
500
0
dchain
(i)
55.9017
650
0
FIG. 7: The mean FPT Ti,j as a function of distance dchain between initial position (i, j) and the DNA chain. The diffusion
constant D1 used in calculations of each row is D1 = 0.1, 1, 10 (from the top down), and D2 used in each column is D2 = 1, 10, 100
(from left to right).
fined similarly.
From all plots in Figs. 6 and 7, as well as further detailed analyses of the calculation results, we conclude that the
mean FPT T (dtarget, dchain) increases with both distance dtarget and distance dchain roughly. For large values of D2
while small values of D1, the influence of distance dchain is negligible, which implies that the one-dimensional sliding
on DNA chain has almost no contribution to increase the search efficiency of protein.
V. THE DIRECT BINDING PROBABILITY P direct
12
In this section, we will discuss the dependence of direct binding probability P direct on diffusion rates D1 and D2,
see Section ?? of the Supplementary Material for the method used in this study to get P direct. As has been defined
before, P direct describes the probability that protein reaches its target binding site through its adjacent position in
cell space but not its nearest neighbor sites on DNA chain. For convenience, we define P direct
i,j
as the direct binding
probability of a protein initiated at position (i, j) in the cell space, and then P direct is obtained as the average of
P direct
i,j
,
P direct = PM
i=1 PN
j=1 P direct
M N
i,j
.
(13)
For given values of L, kb, kd and kt, the numerical results of P direct are plotted in Fig. 5(d) for different values of
diffusion rates D1 and D2. We found that P direct increases with two-dimensional diffusion rate D2 while decreases
with one-dimensional diffusion rate D1. Which implies that for large D2 but small D1, protein will reach its target
binding locus mainly through two-dimensional diffusion, and the contribution of one-dimensional sliding along DNA
chain is not significant. One the contrary, for small D2 but large D1, one-dimensional sliding along DNA plays main
role on the target search process of protein. One can also find from Fig. 5(d) that the limit value of P direct for
D2 → ∞ depends on D1, while the limit value for D2 → 0 does not. The properties of typical examples P direct
,
1,(N +1)/2 are similar as those of P direct, see Figs. ??(d), ??(d), and ??(d). Nevertheless, the plots
P direct
(M+1)/2,1 and P direct
in Fig. ??(d) show that the behavior of P direct
(M+1)/2,(N +1)/2 is different. It increases to value one monotonically as
two-dimensional diffusion rate D2 decreases, but still decreases with one-dimensional diffusion rate D1. The behavior
of average value P direct is similar as those of P direct
with (i, j) near the boundary of the two-dimensional cell space.
1,1
i,j
As mentioned in Section IV A, this is due to the effect of high dimension that most of the cell positions are closer to
the cell boundary than to the cell center, which is assumed to be the position of the target binding locus of DNA
chain in this study.
VI. CONCLUDING AND REMARKS
In this study, a simple chain-space coupled model is employed to discuss the search process of a protein molecule in
cell space to its target locus on DNA chain. The protein molecule may be RNA polymerase or transcription factor,
and the binding of it to certain locus of DNA is essential to regulate the expression of gene. How these protein
molecules find their corresponding target loci on DNA remains unclear. In our study, the DNA chain is simplified to
be a one-dimensional lattice, and the cell space is simplified to be a rectangular lattice. The mean first-passage time
(FPT) is chosen as one criterion to evaluate the search efficiency.
Our results show that there exists one critical value of the one-dimensional diffusion rate D1. With which the search
efficiency of a protein molecule is almost independent of the frequency and time that the protein molecule slides along
DNA chain. If the value of D1 is larger than this critical value, the one-dimensional sliding along DNA chain will
be helpful to increase the search efficiency of protein, while for D1 lower than this critical value, sliding along DNA
chain will have no contribution to the search process.
Meanwhile, this study found that the search efficiency of protein increases first and then decreases with the length
13
of DNA chain. Which implies that only bindings to sites of DNA chain near the target locus can help to increase
the search efficiency. This is consistent with the claim given in [39] that most proteins bind the target locus of DNA
chain without long-range one-dimensional sliding.
The probability P direct that a protein molecule reaches its target locus through direct binding from nearby position
in cell space (but not adjacent sites on DNA chain) is also discussed in this study. Our results show that P direct
increases with two-dimensional diffusion rate D2 while decreases with one-dimensional diffusion rate D1. Therefore,
for high values of D2 but low values of D1, protein molecule will reach its target locus mainly through diffusion in
cell space. One the contrary, for low values of D2 but high values of D1, protein molecule will reach its target mainly
through one-dimensional sliding along DNA chain.
The results obtained in this study will help to understand the target search process in cells during gene expression.
Which show that how RNA polymerase or transcription factor reaches their binding sites on DNA, i.e. mainly
through high dimensional diffusion or with the help of one-dimensional sliding along DNA, depends on the detailed
environment of cells (such as diffusion constants) as well as the properties of RNA polymerase or transcription factor
(binding/unbinding rate to/from DNA, and etc).
Y.Z. designed the research; J.L. and Y.Z. performed the research and wrote the paper, J.L. wrote programs.
Author Contributions
This study was supported by the Natural Science Foundation of China (Grant No. 11271083).
Acknowledgments
[1] B. Alberts, A. Johnson, J. Lewis, M. Raff, K. Roberts, and P. Walter, Molecular Biology of the Cell (Garland Science,
2007), 5th ed.
[2] H. Buc and W. R. McCluret, Biochemistry 24, 2712 (1985).
[3] P. L. Dehaseth, M. L. Zupancic, and M. T. R. Jr., Journal of Bacteriology 180, 3019 (1998).
[4] P. R. Jensen and K. Hammer, Biotechnology and Bioengineering 58, 191 (1998).
[5] A. Sanchez, H. G. Garcia, D. Jones, R. Phillips, and J. Kondev, PLoS Computational Biology 7, e1001100 (2011).
[6] A. Vannini and P. Cramer, Molecular Cell 45, 439 (2012).
[7] R. M. Saecker, M. T. R. Jr., and P. L. deHaseth, Journal of Molecular Biology 412, 754 (2011).
[8] O. G. Berg and P. H. .von Hippel, Journal of Molecular Biology 193, 723 (1987).
[9] H. Alper, C. Fischer, E. Nevoigt, and G. Stephanopoulos, Proceedings of the National Academy of Sciences of the United
States of America 102, 12678 (2005).
[10] R. S. Cox, M. G. Surette, and M. B. Elowitz, Molecular Systems Biology 3, 1483 (2007).
[11] V. A. Rhodius and V. K. Mutalik, Proceedings of the National Academy of Sciences of the United States of America 107,
2854 (2010).
[12] D. J. Lee, S. D. Minchin, and S. J. W. Busby, Annual Review of Microbiology 66, 125 (2012).
[13] J. Elf, G. W. Li, and X. S. Xie, Science 316, 1191 (2007).
[14] A. Tafvizi, F. Huang, A. R. Fersht, L. A. Mirny, and A. M. van Oijen, Proceedings of the National Academy of Sciences
108, 563 (2011).
[15] S. E. Halford and J. F. Marko, Nucleic Acids Research 32, 3040 (2004).
[16] A. B. Kolomeisky, Physical Chemistry Chemical Physics 13, 2088 (2011).
[17] L. Mirny, M. Slutsky, Z. Wunderlich, A. Tafvizi, J. Leith, and A. Kosmrlj, Journal of Physics A: Mathematical and
14
Theoretical 42, 434013 (2009).
[18] J. H. Roe, R. R. Burgess, and M. T. R. Jr, Journal of Molecular Biology 176, 495 (1984).
[19] L. J. Friedman and J. Gelles, Cell 148, 679 (2012).
[20] H. Bujard, Promoters: Structure and Function (Praeger, New York, 1982).
[21] O. G. Berg, R. B. Winter, and P. H. V. Hippel, Biochemistry 20, 6929 (1981).
[22] R. B. Winter and P. H. V. Hippel, Biochemistry 20, 6948 (1981).
[23] S. E. Halford, Biochemical Society Transactions 37, 343 (2009).
[24] A. G. Cherstvy, A. B. Kolomeisky, and A. A. Kornyshev, Journal of Physical Chemistry B 112, 4741 (2008).
[25] H. X. Zhou, Proceedings of the National Academy of Sciences 108, 8651 (2011).
[26] J. Reingruber and D. Holcman, Physical Review E 84, 020901(R) (2011).
[27] Y. von Hansen, R. R. Netz, and M. Hinczewski, Journal of Chemical Physics 132, 135103 (2010).
[28] H. Kabata, O. Kurosawa, I. Arai, M. Washizu, S. A. Margarson, R. E. Glass, and N. Shimamoto, Science 262, 1561 (1993).
[29] C. S. Park, Z. Hillel, and C. W. Wu, Journal of Biological Chemistry 257, 6944 (1982).
[30] C. S. Park, F. Y. Wu, and C. W. Wu, Journal of Biological Chemistry 257, 6950 (1982).
[31] P. Singer and C. W. Wu, Journal of Biological Chemistry 262, 14178 (1987).
[32] P. T. Singer and C. W. Wu, Journal of Biological Chemistry 263, 4208 (1988).
[33] M. Guthold, X. Zhu, C. Rivetti, G. Yang, N. H. Thomson, S. Kasas, H. G. Hansma, B. Smith, P. K. Hansma, and
C. Bustamante, Biophysical Journal 77, 2284 (1999).
[34] Y. Harada, T. Funatsu, K. Murakami, Y. Nonoyama, A. Ishihama, and T. Yanagida, Biophysical Journal 76, 709 (1999).
[35] M. Ricchetti, W. Metzger, and H. Heumann, Proceedings of the National Academy of Sciences 85, 4610 (1988).
[36] K. Sakata-Sogawa and N. Shimamoto, Proceedings of the National Academy of Sciences 101, 14731 (2004).
[37] L. Bai, T. J. Santangelo, and M. D. Wang, Annual Review of Biophysics and Biomolecular Structure 35, 343 (2006).
[38] F. Wang and E. C. Greene, Journal of Molecular Biology 412, 814 (2011).
[39] L. J. Friedman, J. P. Mumm, and J. Gelles, Proceedings of the National Academy of Sciences 110, 9740 (2013).
[40] F. Wang, S. Redding, I. J. Finkelstein, J. Gorman, D. R. Reichman, and E. C. Greene, Nature Structural & Molecular
Biology 20, 174 (2012).
[41] A. B. Kolomeisky and A. Veksler, Journal of Chemical Physics 136, 3304 (2012).
[42] S. Redner, A Guide to First-Passage Processes (Cambridge University Press, New York, 2001).
[43] T. Hastie, R. Tibshirani, and J. Friedman, The Elements of Statistical Learning: Data Mining, Inference, and Prediction,
Springer Series in Statistics (Springer, 2011), 2nd ed.
|
1212.4470 | 2 | 1212 | 2013-08-20T19:25:51 | Noise-aided Logic in an Electronic Analog of Synthetic Genetic Networks | [
"physics.bio-ph",
"q-bio.MN"
] | We report the experimental verification of noise-enhanced logic behaviour in an electronic analog of a synthetic genetic network, composed of two repressors and two constitutive promoters. We observe good agreement between circuit measurements and numerical prediction, with the circuit allowing for robust logic operations in an optimal window of noise. Namely, the input-output characteristics of a logic gate is reproduced faithfully under moderate noise, which is a manifestation of the phenomenon known as Logical Stochastic Resonance. The two dynamical variables in the system yield complementary logic behaviour simultaneously. The system is easily morphed from AND/NAND to OR/NOR logic. | physics.bio-ph | physics |
1
Noise-aided Logic in an Electronic Analog of Synthetic Genetic
Networks
Edward H. Hellen1,∗, Syamal K. Dana2, Jurgen Kurths3, Elizabeth Kehler1, Sudeshna Sinha4
1 Department of Physics and Astronomy, University of North Carolina Greensboro,
Greensboro, NC 27402, USA
2 CSIR-Indian Institute of Chemical Biology, Kolkata 700032, India
3 Potsdam Institute for Climate Impact Research, Telegrafenberg A31 14473 Potsdam
14473 Potsdam, Germany
4 Indian Institute of Science Education and Research (IISER) Mohali, SAS Nagar, Sector
81, Mohali 140 306, Punjab, India
∗ E-mail: [email protected]
Abstract
We report the experimental verification of noise-enhanced logic behaviour in an electronic analog of a
synthetic genetic network, composed of two repressors and two constitutive promoters. We observe good
agreement between circuit measurements and numerical prediction, with the circuit allowing for robust
logic operations in an optimal window of noise. Namely, the input-output characteristics of a logic gate is
reproduced faithfully under moderate noise, which is a manifestation of the phenomenon known as Logical
Stochastic Resonance. The two dynamical variables in the system yield complementary logic behaviour
simultaneously. The system is easily morphed from AND/NAND to OR/NOR logic.
Introduction
Realization of logic functions in different physical systems is one of the key questions that commands
widespread research interest in science and engineering. Universal general-purpose computing devices can
be constructed entirely from NOR/NAND logic gates [1, 2]. It is particularly interesting to investigate if
systems of biological relevance can also yield logic outputs consistent with the truth tables of different
logic functions (see Table 1). Biological systems are capable of stochastic resonance [3 -- 6], a process
in which a small signal is amplified due to the presence of an appropriate level of noise, leading to the
possibility of a biological system performing robust noise-aided logic operations in response to weak input
signals.
A new idea in this direction uses the interplay between noise and nonlinearity constructively to
enhance the robustness of logic operations. Namely, in an optimal window of noise, the input-output
characteristics of a logic gate is reproduced faithfully. This phenomenon is termed Logical Stochastic
Resonance (LSR) [7 -- 12]. Specifically, in LSR we consider the state of a nonlinear system when driven
by input signals, consisting of two randomly streaming square waves. It was observed that the response
of such a system shows a remarkable feature:
in an optimal band of noise, the output of the system,
determined by its state, is a logical combination of the two input signals in accordance with the truth
tables of fundamental logic operations.
An important motivation for further studying LSR stems from an issue that is receiving widespread
attention currently. The number of transistors in an integrated circuit has approximately doubled every
year in accordance with Moore's law. The rapid shrinking of computing platforms with smaller power
supplies has brought with it problems of smaller noise margins and higher error rates. Namely, as
computational devices and platforms continue to shrink in size, we encounter fundamental noise that
cannot be suppressed or eliminated. Hence an understanding of the cooperative behavior between a device
noise-floor and its nonlinearity plays an increasingly crucial role in paving the way for smart computing
devices. In this direction, LSR indicates a way to turn potentially performance degrading noise to assist
2
the desired operation. Further, it is of far reaching interest to obtain analogous behaviour, not merely in
human engineered physical systems, but also in systems of chemical and biological relevance, in order to
explore the information processing capacity of naturally occurring systems where noise is ubiquitous.
Since the idea of LSR was first introduced [7], several systems implementing and displaying LSR
have been found. To begin with, the basic electronic realizations of simple bistable potentials were
reported [7, 8]. Subsequently, noise-aided reprogrammable logic gates have been implemented with noisy
nanomechanical oscillators [12], chemical systems [12] and optical systems [13, 14].
Most recently, in the context of biological systems, theoretical ideas have been proposed [15 -- 18] on the
implementation of LSR in a synthetic genetic network [19]. Now, in this work, we will provide experimental
realizations of these ideas in an electronic analog of a noisy synthetic gene network. Specifically then, we
will investigate the possibility of obtaining reliable logic outputs, and explicitly demonstrate the pivotal role
of noise in the optimization of the logic performance in this circuit. Further, we will show that the system
is easily changed from AND/NAND logic to OR/NOR demonstrating potential for re-programmability
[15,16]. Our results will thus provide verification and further understanding of noise aided logic in systems
that are of considerable importance in biology.
Since understanding the intracellular processes in a network of interacting biomolecules is difficult, an
alternative approach has been started recently [20], to design artificial genetic networks to derive desired
functional behaviors. One important early design is a clock using three genes inhibiting each other
in a cyclic order [21]. Taking into account the standard chemical kinetics for expression, degradation
and inhibition, a dynamical system model was proposed where the repressor-protein concentrations and
mRNA concentrations were expressed as dynamical variables. Another design is a synthetic genetic
toggle-switch network [22] whose potential for noise-aided logic operation is investigated here.
The significance of using both numerical simulation and electronic circuits to model a potential syn-
thetic genetic network is two-fold. Firstly, the numerical and circuit methods provide different imperfect
models of a potential biological system. Agreement between these two models indicates robustness in the
system and therefore greater likelihood that the same behavior could be realized in the proposed biological
system. The biological system is generally much more difficult to construct, and therefore investigating
proposed networks in simpler systems is prudent. Secondly, modeling with stochastic differential equa-
tions is nontrivial compared to ordinary differential equations, so that the addition of experimental
measurements from a physical system such as an analog circuit provides valuable verification. Thus the
circuit is an additional tool for investigating potentially interesting biological networks in the presence of
noise.
Here we use two repressors and constitutive promoters as our model system for implementing logic
functions. First we describe the synthetic gene network model below, and define what constitutes logic
inputs and logic outputs in this system. We then go on to present the electronic analog of the system
followed by a comparison of numerical simulation and experimental measurement.
Methods
We begin with a short description of the general principle of LSR. Consider a general nonlinear dynamic
system, given by
dx
dt
= F (x) + I + Dη(t)
(1)
where F (x) is a generic nonlinear function which has or nearly has two distinct stable energy wells. I is
a low amplitude input signal and η(t) is an additive zero-mean Gaussian noise with unit variance with
D being the noise strength.
We achieve a logical input-output correspondence with such a system by encoding N inputs in N
square waves. Specifically, for two logic inputs, we drive the system with a low amplitude signal I, taken
3
to be the sum of two pulse trains: I1 + I2, where I1 and I2 encode the two logic inputs. Now the logic
inputs can be either 0 or 1, giving rise to 4 distinct logic input sets (I1, I2): (0,0), (0,1), (1,0) and (1,1).
Since the input sets (0,1) and (1,0) give rise to the same I, the input signal generated by adding two
independent input signals is a 3-level aperiodic waveform.
The output of the system is determined by its state. For instance, for a bistable system with wells
at x = x1 and x2, the output can be considered a logical 1 if it is in the well at x1, and logical 0 if it is
in x2. If we consider the opposite assignment, namely logical 1 if the state is in well x2 and logical 0 if
the state is in well x1, we obtain a complementary logic operation. Specifically we can have an output
determination threshold x∗, located near the barrier between the wells, and the logical outputs are then
simply given by the state being greater than or less than x∗. It is possible that the input I induces the
appearance of the second energy well if it was not already there.
The central result of LSR is as follows: for a given set of inputs (I1, I2), a logical output, in accordance
with the truth tables of the basic logic operations, is consistently obtained only in an optimal window
of noise. Namely, under very small or very large noise the system does not yield reliable logic outputs,
while in a band of moderate noise it produces the desired output.
Synthetic Genetic Network Model and Logic Operation
We consider the previously used variation [15] of the genetic toggle switch model comprised of two genes
inhibiting each other [22]. The concentrations of the two expressed proteins are x and y, and their rates
of change are:
dx
dt′ =
dy
dt′ =
α1
1 + yn − β1x + g1 + Dη(t′)
α2
1 + xn − β2y + g2 + I1 + I2 + Dξ(t′)
(2)
(3)
where β1, β2 are the rates of decay of each expressed protein and n is the Hill coefficient. The α1,
α2 describe the maximum expression rates in absence of inhibitor and they are used here as tunable
parameters. In the original model g1 and g2 represent the basal synthesis rates of the promoters [23],
however we use them as constant bias. The additive noise has strength D and η and ξ are chosen from unit
variance zero mean Gaussian distributions. Such an additive noise source alters the background repressor
production and represents the inherent stochasticity of biochemical processes such as transcription and
translation, and the fluctuations in the concentration of a regulatory protein. I1 and I2 are two low
amplitude inputs providing independent parallel production pathways of repressor y. The t′ indicates
dimensionless time.
The system above may have two stable configurations in the xy-plane: one state has a high value of
x (xu) and a low value of y (yl); the other state has a low value of x (xl) and a high value of y (yu).
That is, the two dimensional potential underlying this system has two wells, (xu, yl) and (xl, yu), in the
xy-plane. Varying the parameters changes the depth and position of these wells, and also determines
whether there are one or two wells. For example, Fig. 1 shows that for the case (gi, Ii, D) = 0 the system
in Eqs. (2)-(3) is bistable and therefore has two stable wells only when β1 is close to 1.
Encoding Inputs: Here the low amplitude input signal is I = I1 + I2, with I1/I2 equal to ION
(ION > 0) if the logic input is 1, and I1/I2 being 0 if the logic input is 0. So we have:
(i) I1 + I2 = 0 corresponds to logic input set (0, 0)
(ii) I1 + I2 = ION corresponds to logic input sets (0, 1)/(1, 0)
(iii) I1 + I2 = 2ION corresponds to logic input set (1, 1)
Output: The outputs of the system are determined by the level of the dynamical variables x(t) and
y(t). For instance the output can be considered a logical 1 if the state is at the high level, and logical 0
if it is at the lower level. That is:
4
(i) If x < x∗, then Logic Output is 0
(ii) If x > x∗, then Logic Output is 1
Here x∗ is the output determination threshold that lies between the two states, e.g., at the position of
the barrier between the wells. The results presented here are not sensitive to the specific value of x∗.
Specifically, in this work, we consider the logic output to be 1 when the state is close to the upper
well, and 0 when the state is close to the lower well, for both x and y variables. So when the system
switches wells, the output is "flipped" or "toggled".
The model in Eqs. (2)-(3) is based on the synthetic genetic toggle switch previously expressed in E.
coli [22]. Parameter values used in [22] correspond here to αi = (15.6, 156), n = (1, 2.5), and βi = 1 in
Eqs. (2)-(3). By comparison, here we use αi = 1.78, n = 2.4, β1 = 0.9, and β2 = 1. The bifurcation
diagram in Fig. 1 indicates that these parameter values, along with (gi, Ii, D) = 0, result in a system with
a single stable well at x ≈ 1.8, y ≈ 0.35. A non-zero input I can then "shift" the bifurcation diagram of
Fig. 1 so that there is a stable state with low-x, high-y at β1 = 0.9.
Circuit Realization
The circuit of a single inhibitory gene [24, 25] is shown in Fig. 2. The transistor current represents
the rate of gene expression and the voltage Vout represents the concentration of expressed protein. Vin
represents the concentration of repressor, and the Vcth adjusts the affinity of the repressor binding to
the gene's DNA. The Hill function inhibition in Eqs. (2)-(3) is accounted for by the dependence of the
transistor current on repressor concentration voltage Vin. The synthetic genetic network shown in Fig. 3
is comprised of two individual gene circuits connected in a loop, each inhibiting the other. For the model
in Eqs. (2)-(3), the encoding inputs I1 and I2 add to production of y which is accounted for in Fig. 3 by
the two logic-driven transistors sourcing current to Vy. Initially parameters g1 and g2 are taken to be
zero.
The circuit equations are obtained by applying Kirchoff's laws to Vx and Vy, the voltages across the
capacitors in Fig. 3 [24, 25]. Multiplying both equations by Ry results in equations for Vx and Vy;
RyC
RyC
dVx
dt
dVy
dt
= −
Ry
Rx
Vx +
Ry
Rx
Vx noise + Ryit
= −Vy + Vy noise + Ryit + Ryi1 + Ryi2
(4)
(5)
where it are the Fig. 2 transistor currents for each gene, and i1 and i2 are currents from the logic
train transistors in Fig. 3. A noise generation circuit shown in Fig. 4 based on breakdown of a reverse
biased base-emitter junction produces noise Vnoise with zero mean and variable amplitude. We use a well
regulated supply for the noise circuit to avoid adding AC signals from the building's electrical system
into the noise. Two of these noise circuits are used to supply noisy voltages to each individual gene at
locations indicated in Fig. 3.
The connections between model parameters (αi, βi, n, Ii, gi, D) and circuit parameters are presented
in this section using relevant numerical values, with derivations of these connections given in the next
section. Readers may go directly to Results and Discussion without loss of continuity. The connections
are found by relating circuit Eqs. (4)-(5) to the model Eqs. (2)-(3) and by adjusting the dependence
of the transistor current it on Vin in Fig. 2 to match the Hill function inhibition. The dimensionless
state-variables (x, y) in Eqs. (2)-(3) are related to voltages Vx and Vy by:
x =
Vx
Vth
, y =
Vy
Vth
where Vth corresponds to the repressor's half-maximal inhibition binding constant Ki. The maximal
expression rate is
α1 = α2 =
= 1.78
(6)
imaxR
Vth
=
1.35
0.76
5
where the voltage imaxR is (3mA)(0.45kΩ) = 1.35V and Vth = 0.76V . Protein decay rates are β1 =
Ry
Rx
520 = 0.90, and β2 = 1. The Hill coefficient n comes from
= 470
nα = 1.17G1G−2
(7)
where G1 = −1.1 and G−2 = −3.3 are closed loop gains of U1 and U2 in Fig. 2 resulting in n =
(1.17)(1.1)(3.3)/1.78 = 2.4. The characteristic time is RyC = (470Ω)(2 × 10−8f ) = 9.4µs, so the
dimensionless time is t′ = t
Ry C .
The high value ION of the encoding signals I1, I2 is given by
ION =
Ry × (1V olt)
RI Vth
=
470
RI (0.76)
=
618
RI
(8)
where ION is changed by varying RI in Fig. 3, i.e. RI = 10kΩ gives ION = 0.062. A non-zero value
of g2 in Eq. 3 is achieved in the circuit by including a third current-sourcing transistor in Fig. 3 in the
same way as the two encoding signal transistors, but with the emitter resistor labelled Rg and the input
grounded so that the transistor provides a constant current ig = 1V olt/Rg. g2 is changed by varying Rg
in the same way RI controls ION . The non-zero g2 adds a term Ryig to Eq. 5, where
g2 =
Ryig
Vth
=
470
Rg0.76
=
618
Rg
.
(9)
Rg = 10kΩ gives g2 = 0.062.
The Vnoise terms in Eqs. (4)-(5) approximate white noise voltages. Each Vnoise is characterized
by its measured rms value VN rms and bandwidth. VN rms is controlled by changing the gain via the
potentiometer in Fig. 4. Noise strength D in Eqs. (2)-(3) is given by
D =
VN rms
Vth√γ RC fc1
=
(10)
VN rms
0.76pγ(9.4µs)(1.5M Hz)
where fc1 = 1.5 MHz is the cut-off frequency for the amplifier in Fig. 4 and γ decreases from π/2 at low
gains to π/4 at high gain (when the potentiometer is set to 20kΩ in Fig. 4).
Circuit Analysis and Simulation
Here we describe the circuit analysis and derive the connections between the model parameters used in
Eqs. (2)-(3) and the circuit parameters. Further details are given in Refs. [24, 25].
The single gene circuit in Fig. 2 is designed to reproduce the Hill function inhibition in Eqs. (2)-(3).
The op-amp U1 is configured as a subtraction amplifier with gain G1 = − 11
10 = −1.1. Replication of the
Hill function behavior is achieved by allowing saturation of the output of the op-amp U2 and by having
different unsaturated gains G+2 and G−2 for Vin > Vcth and Vin < Vcth, respectively, due to the diodes
in the feedback for U2. G−2 is the gain of U2 when its output goes negative, in which case the diodes
are not conducting, and therefore G−2 = − 3.3
1.0 = −3.3. G+2 is a diminishing gain when the output
of U2 becomes increasingly positive causing the diodes to go into conduction. An increasing repressor
concentration corresponds to Vin surpassing Vcth which causes the unsaturated output at U2 to change
from a negative voltage of G1G−2(Vin − Vcth) to a positive voltage G1G+2(Vin − Vcth). The increasing
voltage at the output of U2 turns the transistor off (it → 0) which corresponds to complete inhibition of
protein expression. Maximal protein expression α1,2 in Eqs. (2)-(3) corresponds to the maximum value
of it, designated imax. it = imax occurs when Vin = 0 (no repressor) because the output of U2 saturates
at V−sat = −3.5 V (for the LF412 op-amp supplied by ±5 V), resulting in a 0.65 V drop across the 220Ω
and therefore imax = 3 mA.
6
Circuit parameters for α and β are found by using Vth to convert Eqs. (4)-(5) to a dimensionless
form for comparison to Eqs. (2)-(3). The 0.45kΩ used for R in Eq. (6) comes from the Ry = 470Ω being
nearly in parallel with the resistance (10 kΩ) at the input to the subtraction amplifier U1 for gene-x.
The relation for Hill coefficient n is found by adjusting the dependence of it on Vin to match the slope of
the Hill function 1/(1 + xn) at x = 1 resulting in the relation [25]:
−f VthG1G−2
f (5 − V−sat) − 0.6
= −n
4
.
(11)
In Fig. 2 the voltage divider fraction f = 0.4/2.6 = 0.154 and V−sat = −3.5 V. Using Eq. (6) in Eq. (11)
yields Eq. (7).
Parameter Vth's correspondence to binding affinity of repressor to DNA is seen by noting that Eqs.
(2)-(3) are dimensionless, meaning that in the process of going from chemical kinetic equations to Eqs.
(2)-(3) the maximal expression rate α has been scaled by the repressor's inhibition binding constant Ki,
and by a mRNA degradation rate [21, 25]. From Eq. (6) it follows that Vth is proportional to Ki since α
is inversely proportional to Ki due to the scaling. To find the relation between Vcth in Fig. 2 and Vth we
note that the Hill function equals 0.5 when x = 1. Therefore it must be half its maximum value when
Vin = Vth which gives [25]
it
f (5 − G1G−2(Vth − Vcth)) − 0.55
=
= 0.5.
imax
Solving gives Vcth ≈ Vth + 1
. Using n = 2.4 and α = 1.78 in Eqs. (6)-(7) gives: G1G−2 = 3.65,
satisfied by G1 = −1.1 and G−2 = −3.3; Vth = 1.35/1.78 = 0.76 V; and Vcth = 0.76 + 1/3.65 = 1.03 V.
Comparing Eqs. (3) and (5) shows that the encoding signals I1, I2 are related to the transistor currents
i1, i2 by:
f (5 − V−sat) − 0.6
G1G−2
I1,2 =
Ryi1,2
Vth
.
(12)
The encoding currents i1,2 take on two possible values depending on whether their logic train input in
Fig. 3 is high or low. When the input is high (> 4V) the transistor is off so the current is zero. When
the input is zero, the voltage divider consisting of the 4.7kΩ and 2.2kΩ produces 1Volt across the RI
connected to the emitter of the pnp transistor creating current i1,2 = (1V olt)/RI . Equation (12) then
gives Eq. (8) for ION . Results of an analysis for a non-zero value of parameter g2 are the same as for
encoding signals I1, I2 and currents i1, i2 because g2's sourcing transistor is set up in the same way as
the transistors in Fig. 3 for the encoding signals. Thus the non-zero value of g2 is Eq. (9).
Here we show how to use simulations to predict the circuit results. In the process we find Eq. (10),
the connection between circuit parameters and the noise strength D in Eqs. (2)-(3). A standard Euler-
Maruyama simulation of Eq. (2) is
xi+1 − xi = (cid:18) α1
1 + yn
i − β1xi + g1(cid:19) ∆t′ + D√∆t′ ηi (0, 1)
(13)
where ηi(0, 1) is a unit variance zero mean normal random distribution. The noise circuit in Fig. 4
produces a measurable rms voltage VN rms consisting of contributions from all the frequency components
present in the noise. The variance of the noise is the integral of its spectral density function SD(f ) over
frequency, and is obtained from a measurement of VN rms;
V 2
N rms = Z SD(f ) df .
(14)
Idealized white noise assumed in Eqs. (2)-(3) has a SD(f ) which is uniform over an infinite bandwidth.
However for the real noise from the 2-stage noise amplifier circuit in Fig. 4 each op-amp's gain-bandwidth
product produces a high frequency cut-off, fc1 and fc2. The resulting SD(f ) has the form
SD(f ) =
SD0
(cid:16)1 + (f /fc1)2(cid:17)(cid:16)1 + (f /fc2)2(cid:17)
7
(15)
where SD0 is a constant related to the strength of the noise. The OPA228 op-amp has a gain-bandwidth
product 33 MHz, therefore the first stage in Fig. 4 with fixed gain 22× has cut-off, fc1 = 33/22 = 1.5
MHz. The second stage's cut-off fc2 varies from 33 to 1.5 MHz depending on the gain setting determined
by the potentiometer in the feedback of the second stage op-amp.
The integration in Eqs. (14)-(15) gives
V 2
N rms = Z ∞
0
SD0
(cid:16)1 + (f /fc1)2(cid:17)(cid:16)1 + (f /fc2)2(cid:17)
df =
fc1fc2
fc1 + fc2
π
2
SD0.
(16)
There are two limiting cases for the integral: for small gains fc2 ≈ 33 MHz >> fc1 giving (π/2)fc1SD0;
and for large gain (potentiometer → 20kΩ in Fig. 4) fc2 ≈ fc1 giving (π/4)fc1SD0. Thus the integral in
Eq. (16) is γfc1SD0 where fc1 = 1.5 MHz and γ varies from π/2 for small noise to π/4 for large noise.
For frequencies within the noise bandwidth (< 1.5 MHz) the amplitude spectral density ASD (units Volt
Hz-1/2) has a constant value given by
ASD0 = pSD0 =
VN rms√γfc1
(17)
ASD0 is a good approximation of the white noise strength provided that the Fig. 3 genetic circuit's
characteristic response rate 1/(RC) is much less than the noise bandwidth, meaning that 1/(RC) <<
2πfc1. This condition is ensured since the RC used here is (470Ω)(0.02µf ) = 9.4µs, and the bandwidth
of the noise is fc1 ≈ 1.5 MHz.
Figure 5 shows the measured frequency content from the noise circuit when the potentiometer at the
second stage is set for gain 11× producing VN rms = 0.90 V and fc2 = 33/11 = 3 MHz. In this case
fc2 = 2fc1 and γ = π/3. Figure 5 shows that the frequency content is relatively flat out to the cut-
off near 1.5 MHz and therefore is a good approximation to white noise for the genetic network circuit.
VN rms = 0.90 V is on the high end of the noise amplitudes used here.
The circuit Eqs. (4)-(5) which need to be simulated are of the form
dV
dt
=
f (V )
RC
+
Vnoise
RC
(18)
where Vnoise approximates a white noise voltage. Vnoise is characterized by its measured rms value
VN rms and bandwidth, and Eq. (17) gives the noise's amplitude spectral density. The Euler-Maruyama
simulation for Eq. 18 is
(19)
(20)
Vi+1 − Vi =
f (Vi)
RC
∆t +
ASD0
RC
√∆t ηi (0, 1) .
Using dimensionless time ∆t′ = ∆t/(RC) and the measured noise amplitude VN rms gives
Vi+1 − Vi = f (Vi)∆t′ +
VN rms
√γ RC fc1
√∆t′ ηi (0, 1)
Normalizing by the voltage scale Vth puts Eq. (20) in the form of Eq. (13) and gives the connection
between circuit parameters and dimensionless noise amplitude D shown in Eq. (10). For example, using
Vth = 0.76V , RC = (470Ω)(0.02µf ), γ = π/2, and fc1 = 1.5 MHz gives D = 0.28VN rms.
8
Results and Discussion
Figure 6 shows simulations and circuit measurements for three values of noise using parameters n = 2.4,
α = 1.78, β1 = 0.90, β2 = 1, ION = 0.067, and g1 = g2 = 0. It is apparent that for an optimal noise
level (Fig. 6b) the circuit indeed performs the logic AND/NAND function, and that for the smaller (Fig.
6a) and larger (Fig. 6c) noise values faithful logic response is lost. At the low noise (Fig. 6a) the outputs
sometimes fail to respond to the 0 to 1 transition from the AND of I1, I2, and for the 1 to 0 transition
the outputs often wait until both inputs go low before responding thereby causing a delayed response. In
Fig. 6b the responses are quick for both the up and down transition. Figure 6c shows that at the high
noise level the responses are again quick as in 6b, but the outputs also make erroneous transitions. The
circuit behavior is seen to be in agreement with the simulations of Eqs. (2)-(3).
In order to investigate the range of noise strengths which produce accurate logic response, and to find
optimal values for the amplitude ION of the small signal inputs I1,2 we define an accuracy measure a,
a = aL × aH
(21)
where aL is the percent of time the x, y outputs are correct when the AND operation of I1 and I2 is
low, and aH is the percentage of time correct when the AND operation is high. This definition has
the property that when the x, y outputs do not respond at all, then a = 0 since aH = 0 even though
aL = 1. The expectation then is that for no noise there should be no stochastic resonance response so
that a = 0, and that for extreme noise each accuracy would approach 50% so that a → 0.25. If x, y
respond immediately with no mistakes then a = 1. Figure 7 shows accuracy a for simulations and circuit
measurements as a function of noise strength for different values of encoding amplitude ION .
Figure 7a shows that for a small value of encoding amplitude, ION = 0.051 the network is not able
to give a faithful logic response at any noise level. The response at low noise and at high noise are as
predicted, a = 0 and a → 0.25, respectively, but the peak of the window of stochastic resonance response
is well below 1. Figure 7b shows a window of noise providing faithful response for ION = 0.067. The
reason that the accuracy a is slightly below 1 in the window is that the x, y outputs do not respond
immediately to the AND/NAND transitions. This time lag causes the percent of time with incorrect
response to be non-zero and therefore aH and aL are less than 1. In principle an allowance for a time lag
could be included in the calculation of a if it were deemed necessary. However, such an added complication
would not make the noise window any more apparent than it already is in Fig. 7. Figure 7c shows that at
a high value, ION = 0.10, the outputs respond even with no noise, and the addition of noise only creates
more errors. The relative shift between the simulation and circuit accuracies is due to assumptions made
about the noise spectral density function and the integration in Eq. (16) leading to Eq. (20) which gives
the connection between the measured noise amplitude VN rms and dimensionless noise D. In the idealized
case Eq. (16) finds γ ranges from π/2 to π/4 in Eq. (20), with π/2 being appropriate for the noise levels
used in Fig. 6. Adjusting the value of γ can eliminate the relative shift, but there is nothing to be gained
since the appearance of an optimal noise window for LSR at an appropriate value of ION is already
apparent.
One can also reconfigure the system to another set of logic functions, namely the fundamental
OR/NOR logic, by simply including a non-zero value for g2. The parameter g2 effectively changes
the relative position and depth of the wells of the bistable system, allowing the response to morph from
AND/NAND to OR/NOR. For instance changing g2 from 0 to 0.062 (with all other parameters un-
changed) changes the bifurcation diagram from that in Fig. 1 to Fig. 8 showing that the system is now
bistable for β1 = 0.9. The result is that the system displays a clear OR and the complementary NOR
response as shown by the simulation and circuit results in Fig. 9. The low noise case Fig. 9a shows that
at this low noise level the system usually fails to respond. Figure 9a also shows the resting states are
reversed from the g2 = 0 case in Fig. 6a. Figure 9b shows the OR/NOR response at a noise value within
the window for LSR, and Fig. 6c shows errors when the noise is too large.
9
In summary, our results show that the dynamics of the two variables x and y with g2 = 0, mirror
AND and the complementary NAND gate characteristics. Further, when g2 6= 0, we obtain a clearly
defined OR/NOR gate. Since x is low when y is high and vice-versa, the dynamics of the two variables
always yield complementary logical outputs, simultaneously. That is, if x(t) operates as NAND/NOR,
y(t) will give AND/OR.
These results extend the scope and indicate the generality of the recently observed phenomena of
Logical Stochastic Resonance through experimental verifications. Further, these observations may provide
an understanding of the information processing capacity of synthetic genetic networks, with noise aiding
logic patterns. It also may have potential applications in the design of biologically inspired gates with
added capacity of reconfigurability of logic operations.
We have also demonstrated that the electronic circuit provides an additional tool for investigating
dynamics of proposed genetic networks. The circuit measurements are complementary to numerical
simulations, thereby giving indication of the robustness of a particular network design and potential for
successful realization in a biological system.
Thus the results presented in this work suggest new directions in biomolecular computing, and indicate
how robust computation may be occurring at the scale of regulatory and signalling pathways in individual
cells. Design and engineering of such biologically inspired computing systems not only present new
paradigms of computation, but can also potentially enhance our ability to study and control biological
systems [26].
Acknowledgments
S.K.D. acknowledges support by the CSIR Network project GENESIS (BC0123) and the CSIR Emeritus
Scientist scheme.
References
1. Mano MM (1993) Computer system architecture (3. ed.). Pearson Education, I-XIII, 1-524 pp.
2. Bartee T (1991) Computer architecture and logic design. Computer science series. McGraw-Hill.
URL http://books.google.com/books?id=b9dQAAAAMAAJ.
3. Bulsara A, Jacobs E, Zhou T, Moss F, Kiss L (1991) Stochastic resonance in a single neuron model:
Theory and analog simulation. Journal of Theoretical Biology 152: 531 - 555.
4. Levin JE, Miller JP (1996) Broadband neural encoding in the cricket cereal sensory system en-
hanced by stochastic resonance. Nature 380: 165 - 168.
5. Gammaitoni L, Hanggi P, Jung P, Marchesoni F (1998) Stochastic resonance. Rev Mod Phys 70:
223 -- 287.
6. Hanggi P (2002) Stochastic resonance in biology how noise can enhance detection of weak signals
and help improve biological information processing. ChemPhysChem 3: 285 -- 290.
7. Murali K, Sinha S, Ditto WL, Bulsara AR (2009) Reliable logic circuit elements that exploit
nonlinearity in the presence of a noise floor. Phys Rev Lett 102: 104101.
8. Murali K, Rajamohamed I, Sinha S, Ditto WL, Bulsara AR (2009) Realization of reliable and
flexible logic gates using noisy nonlinear circuits. Applied Physics Letters 95: 194102.
9. Guerra DN, Bulsara AR, Ditto WL, Sinha S, Murali K, et al. (2010) A noise-assisted repro-
grammable nanomechanical logic gate. Nano Letters 10: 1168-1171.
10
10. Worschech L, Hartmann F, Kim TY, Hofling S, Kamp M, et al. (2010) Universal and reconfigurable
logic gates in a compact three-terminal resonant tunneling diode. Applied Physics Letters 96:
042112.
11. Fierens P, Ibez S, Perazzo R, Patterson G, Grosz D (2010) A memory device sustained by noise.
Physics Letters A 374: 2207 - 2209.
12. Sinha S, Cruz JM, Buhse T, Parmananda P (2009) Exploiting the effect of noise on a chemical
system to obtain logic gates. EPL (Europhysics Letters) 86: 60003.
13. Singh KP, Sinha S (2011) Enhancement of "logical" responses by noise in a bistable optical system.
Phys Rev E 83: 046219.
14. Perrone S, Vilaseca R, Masoller C (2012) Stochastic logic gate that exploits noise and polarization
bistability in an optically injected vcsel. Opt Express 20: 22692 -- 22699.
15. Ando H, Sinha S, Storni R, Aihara K (2011) Synthetic gene networks as potential flexible parallel
logic gates. EPL (Europhysics Letters) 93: 50001.
16. Dari A, Kia B, Bulsara AR, Ditto W (2011) Creating morphable logic gates using logical stochastic
resonance in an engineered gene network. EPL (Europhysics Letters) 93: 18001.
17. Dari A, Kia B, Wang X, Bulsara AR, Ditto W (2011) Noise-aided computation within a synthetic
gene network through morphable and robust logic gates. Phys Rev E 83: 041909.
18. Sharma PR, Somani P, Shrimali MD (2013) Bio-inspired computation using synthetic genetic
network. Physics Letters A 377: 367 - 369.
19. Hasty J, Isaacs F, Dolnik M, McMillen D, Collins JJ (2001) Designer gene networks: Towards
fundamental cellular control. Chaos: An Interdisciplinary Journal of Nonlinear Science 11: 207-
220.
20. Elowitz M, Lim WA (2010) Build life to understand it. Nature 468: 889-890.
21. Elowitz MB, Leibler S (2000) A synthetic oscillatory network of transcriptional regulators. Nature
403: 335-338.
22. Gardner TS, Cantor CR, Collins JJ (2000) Construction of a genetic toggle switch in escherichia
coli. Nature 403: 339-342.
23. Wang J, Zhang J, Yuan Z, Zhou T (2007) Noise-induced switches in network systems of the genetic
toggle switch. BMC Systems Biology 1: 50.
24. Hellen EH, Volkov E, Kurths J, Dana SK (2011) An electronic analog of synthetic genetic networks.
PLoS ONE 6: e23286.
25. Hellen EH, Dana SK, Zhurov B, Volkov E (2013) Electronic implementation of a repressilator with
quorum sensing feedback. PLoS ONE 8: e62997.
26. Benenson Y (2012) Biomolecular computing systems: principles, progress and potential. Nat Rev
Genet 13: 455-468.
Figure Legends
Tables
11
Figure 1. Bifurcation diagram for state-variables x (solid) and y (dashed). x and y are
complementary outputs, when one is high the other is low. Red (black) indicate stable (unstable) fixed
points. System is bistable for 0.937 < β1 < 1.043. For β1 = 0.9, x is high and y is low. Calculated for
Eqs. (2)-(3) with n = 2.4, αi = 1.78, β2 = 1, Ii = 0, gi = 0, and D = 0.
Figure 2. Circuit for single gene. Inhibitory input at Vin. Expressed protein concentration is
represented by Vout. Ri = 470 Ω for gene-y, 520 Ω for gene-x. Dual op-amp is LF412 supplied by +/-5
V. The pnp transistor is 2N3906. The input noise has a mean of 0 V (gnd) and controllable standard
deviation.
12
Figure 3. Circuit for synthetic genetic network. Encoding inputs are 0 to 5 V pulse trains
creating transistor currents of 0.1 and 0 mA, respectively, for RI = 10kΩ. The x and y gene circuits are
shown in Fig. 2. Each noise input is connected to its own noise circuit shown in Fig. 4.
Figure 4. Noise circuit and its connection to resistor of gene circuit. Source of noise is the
reverse-biased base-emitter junction of the 2N3904 npn transistor on left. OPA2228 dual op-amps
supplied from +/-12 V regulators. OPA2228 has gain-bandwidth product of 33 MHz.
13
Figure 5. Measured frequency spectrum of noise. For 2-stage noise amplifier shown in Fig. 4
with second stage gain 11×. Horizontal scale is 500 kHz/Div, so cursor indicates 1.5 MHz as location of
cut-off frequency. From FFT function on Tektronix TDS 2024B oscilloscope.
Table 1. Logic Table
Input Set (I1,I2) OR AND NOR NAND
(0,0)
(0,1)/(1,0)
(1,1)
0
1
1
0
0
1
1
0
0
1
1
0
Relationship between the two inputs and the output of the fundamental OR, AND, NOR and NAND
logic operations. Note that the four distinct possible input sets (0, 0), (0, 1), (1, 0) and (1, 1) reduce to
three conditions as (0, 1) and (1, 0) are symmetric. Note that any logical circuit can be constructed by
combining the NOR (or the NAND) gates [1, 2].
14
Figure 6. Time series for circuit measurements (upper graph) and simulations (lower) for
different noise strengths showing AND/NAND LSR. Circuit shown in Fig. 3. Simulations are of
Eqns. (2)-(3). Upper red and green traces indicate logic states of the encoding inputs I1,2 and lower
traces show the complementary outputs x(yellow) and y(blue). Panel (b) has noise level within the
optimal range for displaying AND/NAND characteristics. Noise strengths in simulation and in circuit
are: (a) D = 0.043 and VN rms = 200mV , (b) D = 0.107 and VN rms = 500mV , (c) D = 0.15 and
VN rms = 700mV . Voltages and times for the circuit measurements have been converted to
dimensionless quantities as described in the text.
15
Figure 7. Accuracy a of the AND/NAND logic response for simulations (red) and circuit
measurements (blue) as a function of noise strength D and encoding amplitude ION . Noise
strength VN rms for the circuit measurements has been converted to dimensionless strength D using Eq.
(10) with γ = π/2.
Figure 8. Bifurcation diagram for state-variables x (solid) and y (dashed) configured for
OR/NOR. x and y are complementary outputs, when one is high the other is low. Red (black)
indicate stable (unstable) fixed points. System is bistable for 0.870 < β1 < 0.980. Calculated for Eqs.
(2)-(3) with n = 2.4, αi = 1.78, β2 = 1, g1 = 0, g2 = 0.062, Ii = 0, and D = 0.
16
Figure 9. Time series for circuit measurements (upper graph) and simulations (lower) for
different noise strengths showing OR/NOR LSR. Panel (b) has noise level within the optimal
range for displaying OR/NOR characteristics. Noise strengths in simulation and in circuit are: (a)
D = 0.043 and VN rms = 200mV , (b) D = 0.107 and VN rms = 500mV , (c) D = 0.15 and
VN rms = 700mV . Voltages and times for the circuit measurements have been converted to
dimensionless quantities as described in the text.
|
1801.01367 | 1 | 1801 | 2018-01-04T14:29:25 | Similarities between action potentials and acoustic pulses in a van der Waals fluid | [
"physics.bio-ph"
] | An action potential is typically described as a purely electrical change that propagates along the membrane of excitable cells. However, recent experiments have demonstrated that non-linear acoustic pulses that propagate along lipid interfaces and traverse the melting transition, share many similar properties with action potentials. Despite the striking experimental similarities, a comprehensive theoretical study of acoustic pulses in lipid systems is still lacking. Here we demonstrate that an idealized description of an interface near phase transition captures many properties of acoustic pulses in lipid monolayers, as well as action potentials in living cells. The possibility that action potentials may better be described as acoustic pulses in soft interfaces near phase transition is illustrated by the following similar properties: correspondence of time and velocity scales, qualitative pulse shape, sigmoidal response to stimulation amplitude (an `all-or-none' behavior), appearance in multiple observables (particularly, an adiabatic change of temperature), excitation by many types of stimulations, as well as annihilation upon collision. An implication of this work is that crucial functional information of the cell may be overlooked by focusing only on electrical measurements. | physics.bio-ph | physics | Similarities between action potentials and acoustic pulses in a van der
Waals fluid
Matan Mussel∗†‡
Matthias F. Schneider∗
January 8, 2018
Abstract
8
1
0
2
n
a
J
4
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
6
3
1
0
.
1
0
8
1
:
v
i
X
r
a
An action potential is typically described as a purely electrical change that propagates along the membrane
of excitable cells. However, recent experiments have demonstrated that non-linear acoustic pulses that propagate
along lipid interfaces and traverse the melting transition, share many similar properties with action potentials.
Despite the striking experimental similarities, a comprehensive theoretical study of acoustic pulses in lipid systems
is still lacking. Here we demonstrate that an idealized description of an interface near phase transition captures
many properties of acoustic pulses in lipid monolayers, as well as action potentials in living cells. The possibility
that action potentials may better be described as acoustic pulses in soft interfaces near phase transition is illustrated
by the following similar properties: correspondence of time and velocity scales, qualitative pulse shape, sigmoidal
response to stimulation amplitude (an 'all-or-none' behavior), appearance in multiple observables (particularly, an
adiabatic change of temperature), excitation by many types of stimulations, as well as annihilation upon collision.
An implication of this work is that crucial functional information of the cell may be overlooked by focusing only
on electrical measurements.
Keywords: action potential, acoustic pulse, lipid interface, phase transition, van der Waals fluid
1
Introduction
A large category of cells can generate a characteristic
transient change in transmembrane voltage that propa-
gates along the cell membrane in response to suitable
stimuli. These include neurons, myocytes, epithelial
cells, fibroblasts, glia cells, pancreas beta cells as well
as non-specialized cells in corals, plants, fungi, Protozoa
and possibly even bacteria [1–10]. The phenomenon was
first identified in a frog nerve by Emil du Bois-Reymond
in 1843, who called them "action currents", later to be
termed action potentials [1]. Action potentials (APs) are
principally associated with behavioral activities of many
organisms. Therefore, an understanding of the mecha-
nism of APs, as well as their actions and interactions,
constitutes one of the fundamental aspects of biology.
Classically APs are described as a purely electrical
phenomenon, and the non-linear electric response is be-
lieved to be associate with specialized protein compo-
nents [11, 12]. An unavoidable implication of this view
is that the phenomenon should only exist in living cells,
and could not be identified in non-living systems. How-
ever, pulses with similar properties have been observed
in non-living soft systems such as lipids and gels [13,14].
In addition, many experimental facts are neither read-
ily explained nor predicted by the electrical theory of
APs. A partial list includes (1) non-electric aspects that
co-propagate with the electric pulse [15]; (2) the exis-
tence of APs in absence of sodium and potassium in
the intra- and extracellular solutions [16], and in ab-
sence of ion-concentration gradient [17]; and (3) ion-
channel like current fluctuations in the absence of ion
channel proteins [18]. Therefore, the classical picture
not only is not derived from fundamental physical prin-
ciples, it also does not satisfactory describe the observed
phenomenology. Alternative approaches, that treat the
AP on a more physicochemical basis, have been pro-
posed [15,19–21]. A common thread of these approaches
is that electricity is merely one aspect of the pulse, and
it is, therefore, very likely that valuable information is
overlooked due to misunderstanding of its mechanism.
One conjecture, which is the focus of this manuscript,
is that APs are acoustic pulses that propagate along
the lipid interface and cross the phase-transition from
the so called liquid-expanded to the liquid-condensed
phase [19, 20].
Indeed, experimental observations in
lipid monolayers have demonstrated that solitary acous-
tic pulses can be excited, and that they share many
properties with APs. A comparison of key experimental
observations between APs and acoustic pulses in lipid
interfaces is as follows. (1) APs are not purely electrical
waves. Rather, they were identified in multiple observ-
ables – electrical, magnetic, mechanical, optical, as well
as thermal [15, 22, 23]. Acoustic pulses in lipids were
also identified in multiple observables (density, pressure,
electrical, optical and pH) [13, 24–27]. Particularly, the
electric potential difference between the liquid-expanded
and the liquid-condensed phase is ∼100 mV, similar to
the amplitude of APs [24]. (2) The production and ab-
sorption of heat during an AP is adiabatic [28], and an
acoustic description is the natural physical approach for
∗Department of Physics, Technical University of Dortmund, 44227 Dortmund, Germany
†Department of Physics, University of Augsburg, 86159 Augsburg, Germany
‡Correspondence to: [email protected]
1
an adiabatic propagating pulse. (3) The time and veloc-
ity scales of APs vary by 5–6 orders of magnitude be-
tween cells, ∼ 10−3−101 s and ∼ 10−3−102 m/s, respec-
tively [6, 7, 11]. These values are comparable to sound
pulses in lipid systems that traverse the phase transition
(∼ 10−3 − 10−2 s and ∼0.1–1 m/s, respectively) [13, 26].
(4) Excitation of an AP is obtained by many types of
stimulations: electrical, mechanical (pressure, touch, ul-
trasound), optical (by shining light), as well as a change
of temperature (heating or cooling) [1, 29]. Acoustic
pulses in lipids can also be excited by various types of
stimulations (mechanical, electrical and chemical (acid
as well as solvent)) [13, 27, 30, 31]. (5) Even the less in-
tuitive properties of APs, namely threshold behavior (a
sigmoidal response, the so called all-or-none) [32] and
annihilation upon collision [33,34], can be demonstrated
as acoustic behavior near phase transition. Specifically,
stimulation was shown to prompt a sigmoidal response
of density pulses in lipid interfaces near phase transi-
tion [26], and these pulses were demonstrated to annihi-
late upon collision [35].
Despite the striking experimental similarities, a com-
prehensive theoretical study of acoustic pulses that tra-
verse the phase transition in lipid systems is still lack-
ing. Theoretical modeling was initiated by Heimburg
and Jackson, and was focused on solitonic solutions in
a small-amplitude analysis of an acoustic model of the
membrane interface [19]. Two main criticisms of the
model have been an overestimation of the propagation
velocity and the lack of annihilation of pulses. Kap-
pler et al. later demonstrated that the viscous coupling
between the interface and a bulk fluid attenuates the
propagation velocity of interfacial longitudonal waves,
and results in a corrected velocity that agrees with ex-
perimental evidence [36]. Still, some observations are
not captured by these previous works, including the sat-
uration of pulse amplitude at large stimulation ampli-
tudes, annihilation of pulses upon collision, as well as a
thorough analysis of different aspects of the pulses un-
der adiabatic conditions (e.g., density, pressure, temper-
ature and electrical properties).
Herein, we demonstrate that non-linear acoustic
pulses, which annihilate upon collision, arise naturally
from a minimum set of well accepted physical assump-
tions: mass, momentum and energy conservation laws
near a phase transition. Our description relies on rel-
atively few and physically measurable parameters and
variables. Particularly, no assumptions about the molec-
ular constituents and structure, nor any parameter fit,
are required. In spite of its minimalism, the model pro-
vides a rich behavior of non-linear acoustic pulses that
correspond to experimental results in lipid monolayers
as well as to APs in living cells.
2 Model Description
Our purpose is to investigate the properties of isen-
tropic traveling waves which exist near a phase tran-
sition. We therefore consider an idealized ansatz of a
one-dimensional compressible fluid with a van der Waals
constitutive equation, following the works of Felderhof,
Slemrod and Grinfeld [37–39]. While classically a van
der Waals phase transition is between a gas and a liquid,
the familiar pressure-versus-volume curves (isotherms)
qualitatively resemble the melting transition between
the liquid-expanded and the liquid-condensed state in
lipids [40]. The conservation laws of mass, momentum
and energy are written in the Lagrange frame,
∂tw = ¯w∂hv,
∂tv = ¯w∂h(τ1 + τ2),
∂tE = ¯w(cid:2)∂h(τ1v) + k∂2
hθ(cid:3) .
(1)
Here w, v and E denote the specific volume, velocity and
specific total energy of the fluid, respectively. The fluid
density is inversely proportional to the specific volume,
ρ = w−1. The spatial coordinate h (in the Lagrange
frame) is related to the x-axis (in the Euler frame) by
the cumulative mass of fluid particles [41]
wdh,
(2)
(cid:90)
x =
1
¯w
with ¯w a normalization factor that defines the scale of h.
In addition, τ1 and τ2 represent the stresses in the fluid
(see below), k is the coefficient of thermal conductivity
and θ is the fluid temperature.
To allow for a continuous transition between the two
phases, we follow van der Waals' hypothesis that the
energy of a system depends on the gradient of density
(the so called capillarity term) [42, 43]. A similar defini-
tion appears in the Ginzburg-Landau mean field theory
of phase transition as well as the Cahn-Hillard model of
coexistence of two phases [44, 45]. The capillarity coef-
ficient is also known as the gradient-energy coefficient.
The corresponding stress correction to the dynamics of
fluids was developed by Korteweg [42, 46]. Therefore,
the expression for stress includes a capillarity term in
addition to pressure and viscosity. It is defined as
τ1 = −p + ζ∂hv,
τ2 = −C∂2
hw,
(3)
with p the fluid pressure, ζ the dilatational viscosity and
C the capillarity coefficient, treated here as constant for
simplicity.
The system of equations is completed with the van
der Waals constitutive equations, that qualitatively re-
semble the transition between a liquid-expanded and a
liquid-condensed state in lipids (Fig. 1 and Ref. [40])
p =
kBθ
mw − b
E = cvθ − a
m2w
− a
m2w2 ,
v2
2
+
,
(4)
with kB the Boltzmann constant, m the mass of a fluid
particle, a the average attraction between particles, b
the volume exclusion by a fluid particle, and cv the spe-
cific heat capacity [47]. The phase space contains an
unstable region, ∂p/∂w > 0. This region is bounded by
the spinodal curve (marked by S in Fig. 1), which sat-
isfies ∂p/∂w = 0. A slightly larger area, governed by
the Maxwell equal area rule, describes the region where
a coexistence of phases has a lower free energy as com-
pared to a single phase solution (marked by C in Fig. 1).
Both curves meet at the critical point (wc, pc, θc), where
the distinction between the two phases disappears. The
critical point can be expressed by three parameters: m, a
2
Figure 1: Schematic phase diagram in the w–p plane of the van der Waals model. Four different isotherms, along with the Spinodal
(S) and coexistence (C) curves are plotted. The critical point is denoted by a filled circle. LE: liquid-expanded phase, MLE: metastable
liquid-expanded phase, COE: coexistence phase, MLC: metastable liquid-condensed phase, and LC: liquid-condensed phase.
and b (Eq. (S1) in the Supplemental Materials). A de-
tailed list of the 6 variables and 7 parameters of the
model is provided in the Supplemental Materials (Ta-
bles S1–S2).
The critical point and the fluid viscosity were used to
define proper scales (time, length and velocity, respec-
tively)
T ≡ ζ
pc
, L ≡ ζ
√
, U ≡ L
T
=
wcpc.
(5)
(cid:114) wc
pc
These scales were used to derive a dimensionless
form of the model equations (Supplemental Materials,
Eqs. (S3)–(S5)), that depends on only three parameters:
the (dimensionless) heat capacity, thermal conductivity
and capillarity coefficient
cv =
cvθc
pcwc
,
k =
kθc
pcwcζ
,
C =
C
ζ 2 .
(6)
Dimensionless variables and parameters are hereafter
marked with tilde (e.g., x = x/L).
A complete set of all of the material constants is not
known for any single lipid system, nor for a composite
soft interface, for example in biological cells. Therefore,
typical values were estimated from experiments with
DPPC lipids. The critical point of a two-dimensional
DPPC monolayer was identified from isothermal state
diagrams (wc, pc, θc) ∼= (5 · 105 m2/kg, 3 · 10−2 P a ·
m, 315 K) [13, 24, 29, 40]. The order of magnitude of
the membrane viscosity was estimated from measure-
ments of shear viscosity, ζ ∼ 10−3 P a · s · m [48].
The order of magnitude of the heat capacity was evalu-
ated from experiments at constant pressure, cv ∼ cp ∼
103 − 104 J/kg · K [19, 24]. Unfortunately, no direct
measurements of the thermal conductivity and capil-
larity coefficient of lipid monolayers exist. Thermal
conductivity was approximated from the evaluation of
heat conduction in interfacial water in a lipid system,
k ∼ 5 J/s · K [49]. The capillarity coefficient was esti-
mated from measurements of line tension at the phase
boundary, C ∼ 10−27−10−24 kg2/s2 (Supplemental Ma-
terials and Ref. [50]). For typical values of a lipid system,
the proper scales are T≈30 ms, L≈4 m and U≈120 m/s,
and the dimensionless parameters are cv ∼ 101 − 103,
k ∼ 102, and C ∼ 10−21 − 10−18. Although the capil-
larity coefficient is insignificantly small, a non-negligible
value was used for numerical reasons ( C ∼ 1) – to avoid
sharp gradients in the density field. The use of a non-
negligible capillarity coefficient did not show a major
effect on our results (Fig. S1).
The system of equations (1)–(4) was numerically
solved with the Dedalus open-source code [51], which
is based on a pseudo-spectral method. The model was
solved using periodic boundary conditions, and with ho-
mogeneous initial conditions, (w0, p0, θ0) in the liquid-
expanded phase. Excitation of a pulse was obtained by
'injecting' a localized stress (around h0) with an ampli-
tude pexc for a brief time (t0) into the momentum flux;
i.e., adding the following term into the right-hand-side
of the middle Eq. 1
¯w∂h
pexcΘ(t0 − t)e
− (h−h0)2
2λ2
.
(7)
(cid:18)
(cid:19)
Here, Θ is the Heaviside function, and λ is the width of
excitation.
3 Results
We now turn to analyze sound pulses that traverse the
phase transition. Upon excitation, distinct pulses were
obtained when the initial state of the fluid was near
phase transition. Figure 2a (solid line) depicts the shape
of a density pulse as measured at distance x/L = 1 from
the excitation point. The shape of the density pulse was
qualitatively very similar to experimental measurements
in lipid monolayers [13], as well as to the voltage signal
of an AP [32]. Because this issue is of particular rele-
vance to the open debate regarding the physical nature
of cellular excitability, we further discuss the relation be-
tween the interface density and a transmembrane volt-
age measurement in the Discussion section. In contrast,
pulse amplitude was much lower when the initial state
was away from the phase transition (Fig. 2a, dotted-
dashed line). A projection of the closed trajectory in
the w–p plane is shown in Fig. 2b, for the two pulses
3
Figure 2: (a) Density pulse as a function of time, as measured at distance x/L = 1 from the excitation point, with an initial density of
ρ0=0.67 (solid line) and 0.33 (dotted-dashed line). (b) A projection of phase space into the w–p plane (w=ρ−1). Several isotherms are
plotted for reference (shades of grey lines) as well as the coexistence and spinodal curves (dark blue and blue lines, respectively). The
trajectory of the two pulses shown in (a) is plotted in green and purple respectively. (c) Density field of the entire fluid as a function
of space (x-axis) and time (y-axis) with initial density ρ0=0.67. Dashed yellow line marks the solid line solution depicted in (a). Pa-
rameters of the model were (cv, k, C)=(600, 100, 1), additional initial conditions were (v0, θ0)=(0, 0.93), and excitation parameters were
(x0, t0, pexc, λ)=(0, 0.1, 150, 0.088). Numerical calculation was conducted with 4096 grid points, x-domain [−3π/2, 3π/2] and dt = 5· 10−4.
depicted in Fig. 2a (green and purple curves, respec-
tively). If a trajectory does not cross the phase bound-
ary (dotted-dashed line in Fig. 2a and purple curve in
Fig. 2b) the pulse undergoes only a little amplification
in density (ρ (cid:46)1). The change in density is accompanied
by a parallel increase in pressure (p (cid:46)1) and tempera-
ture. Subsequently, the local state decreases in all three
fields (density, pressure and temperature) into a rarefac-
tion state before relaxing back into the initial state. In
contrast, once a pulse traverses the phase transition re-
gion, the density and pressure evolve non-linearly (solid
line in Fig. 2a and green curve in Fig. 2b). At first, a
significant increase in density is obtained (ρ (cid:38)1), with
almost no change in pressure. Subsequently, as the local
state approaches the volume exclusion region, the pres-
sure sharply increases (p (cid:38)1), with only a slight increase
in density. A parallel small increase in temperature also
occurs. Relaxation follows in a reverse order – a sharp
decrease in pressure is followed by a sharp decrease in
density into a rarefaction that relaxes back into the ini-
tial state.
The solution of the density field across the entire fluid
is plotted in Figure 2c. Following an excitation around
x=0 that lasted 0.1T, two localized pulses were gener-
ated, propagating in opposite directions. Their length,
time and velocity scales were ≈0.5L, ≈0.1T, and ≈5U,
respectively. An important observation is that these
pulses did not maintain a constant velocity and shape,
and eventually dispersed; i.e., these solutions are nei-
ther solitons, nor homoclinic orbits. Nonetheless, the
pulses had a distinct shape for a distance of 3–5 times
the pulse width. Variation of the duration or width of
the stimulation (t0 and λ in Eq. (7), respectively) did
not have much influence on the time and length scales
of the pulse (Fig. S2). However, upon increasing the du-
ration of the stimulation (t0), the pulses maintained a
distinct shape for a longer distance (more than 10 times
the pulse width, Fig. S2b). The dashed yellow line in
Fig. 2c marks the solution that was plotted in Fig. 2a
(solid line).
Stimulation by different model variables (veloc-
ity, pressure, temperature or energy) was obtained by
adding the excitation term ( AΘ(t0 − t)e
) to
other model equations (Eqs. (1) or (4)). Figure 3
shows the density pulses that were generated by differ-
ent 'types' of stimulations. The resulting pulses were
qualitatively similar.
− (h−h0 )2
2λ2
By increasing the amplitude of excitation, at a given
initial state, a non-linear (sigmoidal) response of the
density pulse was obtained (Figs. 4a,b). At low am-
plitudes of excitation the pulse response was qualita-
tively similar to previous theoretical results obtained in
a small-amplitude analysis [36]. However, at larger am-
plitudes of excitation the amplitude of the density pulse
saturated at ρ=3ρc. The saturation of density is a direct
result of the exclusion of volume given by the van der
Waals equation (upper Eq. (4) or (S5)). The response
resembles experimental observations in lipid monolay-
ers [26] as well as voltage measurements of APs in living
cells [32].
Figure 4c shows the simultaneous pressure aspect of
the pulse. Pressure increases significantly when the sys-
tem is driven into the liquid-condensed phase. In addi-
4
(a)(b)(c)Figure 3: Excitation by a local injection of (a) velocity (middle of Eq. (1)), (b) pressure (upper Eq. (4)), (c) temperature (lower Eq. (4)),
and (d) energy (lower Eq. (1)). Excitation parameters were ( A, t0) = (150, 0.1), (150, 0.2), (1500, 0.1) and (104, 0.2), respectively. A is the
normalized amplitude of excitation (pexc/pc, pexc/pc, θexc/θc and Eexc/U 2, respectively). Density as a function of time was plotted at
x/L=0.8, 1.0, 0.8 and 1.4, respectively.
tion, during the adiabatic compression and decompres-
sion, the temperature increases and subsequently de-
creases (Fig. 4d). Not surprisingly, the temporal width
of the temperature aspect crucially depends on the ther-
mal conductivity (data not shown).
We proceed now to investigate the question of colli-
sion between longitudonal pulses. While linear (small
amplitude) pulses generally superimpose, these non-
linear waves displayed a rich behavior of interaction, an-
nihilation and in some cases even repulsion. The type
of interaction between pulses strongly depends on the
value of the heat capacity and thermal conductivity of
the fluid. Here we provide an example of annihilation
of two pulses that were excited in a small sized domain.
Figure 5 and Movie S1 show the dynamics of the den-
sity, pressure and temperature fields before, during and
after a collision. The propagation of the pulses is evi-
dent in all three fields, and the collision is characterized
by a localized increase in amplitude, most dramatically
observed in the pressure field.
4 Discussion
The lipid-based interface is a ubiquitous component of
biological cells, and plays a critical role in many cellu-
lar functions. Particularly, the discovery of solitary lon-
gitudinal pulses that travel along lipid interfaces have
stimulated a discussion on the functional role of acous-
tics in biological systems [13, 19, 20, 25–27, 29, 30].
In
this paper we continue this line of research and demon-
strate that an idealized fluid model near phase transi-
tion captures many properties of experimentally mea-
sured acoustic pulses in lipid monolayers as well as APs
in living cells.
Time, length and velocity scales of density pulses
in lipid interfaces. The critical point and the fluid
viscosity specify proper scales of time, length and ve-
locity (Eq. (5)). Specifically, density pulses that tra-
verse the phase transition region scale as ≈0.1T, ≈0.5L
and ≈5U, respectively (Fig. 2). For typical lipid val-
ues these scales correspond to ≈3 ms, ≈2 m, and ≈600
m/s.
In comparison, experimentally measured pulses
in lipid monolayers scale as ≈5–10 ms, ≈1–10 mm and
≈0.1–1 m/s [13, 26], respectively. While the theoreti-
cal time scale agrees with measurements, the length and
velocity scales are overestimated by 2–3 orders of magni-
tude (the same overestimation of velocity was previously
obtained by others [19, 30]). The discrepancy in veloc-
ity and length scales results from our attempt to keep
the model simple, and would disappear by extending
the model to include the non-negligible coupling to the
bulk. It was previously demonstrated that the viscous
coupling between a compressible interface and the bulk
attenuates the velocity and length of acoustic pulses by
ρbηbtp/ρi [36]. Here, ρb is the bulk
density, ηb the bulk viscosity, tp the pulse duration, and
ρi the density of the interface. For a typical lipid inter-
face coupled to bulk water, this results in attenuation
by 2–3 orders of magnitudes, and agrees with experi-
mental observation [30, 36]. Furthermore, these scales
also agree with experimental measurements of APs in
living cells (∼ 10−3 − 101 s and ∼ 10−3 − 102 m/s,
respectively [6, 7, 11]). The variation of 5–6 orders of
magnitudes in time and velocity of APs may be related
to differences in the macroscopic properties of the cell
surface (critical temperature, surface viscosity, relative
state as compared to the critical point), as well as the
bulk (density and viscosity).
a factor of ∼(cid:112)1 +
√
Dispersion of shape and propagation velocity.
According to the classical electrical description, APs are
5
VelocityPressureTemperature Energy(a)(b)(c)(d)Figure 4: (a) Pulse shape at four amplitudes of the excitation, as reflected by the change of density. (b) Non-linear response of the
amplitude of the density pulse (ρmax) to stimulation amplitude (pexc). (c) Pressure and (d) temperature aspects that co-appear with the
density pulse. The first two pressure curves are indistinguishable, as the state did not reach the liquid-condensed phase. Initial density
was ρ0=0.67, and pulse was measured at distance x/L=1.1 from the excitation point. Other parameters appear in the caption of Fig. 2.
pulses that maintain a stable shape and a constant prop-
agation velocity along an 'infinitely' long cell (a homo-
clinic orbit) [11]. Similar characteristics were also sug-
gested by the acoustic soliton model [19]. Surprisingly,
a validation of this hypothesis, by physiological mea-
surements of velocity and shape at more than two sites,
was hardly investigated, presumably due to the small
cellular size. However, recent experiments have showed,
using a multi-electrode array, that the propagation ve-
locity as well as the shape of an AP are not constants
during propagation. Rather, a clear trend of decrease in
velocity and amplitude, as well as an increase in width,
was observed [52, 53]. These findings are in accord with
our results, that do not preserve a constant velocity and
shape, and eventually disperse. We have, nonetheless,
demonstrated that distinct pulses were maintained at
distances up to 1–10 times the length of a pulse (Figs. 2c
and S2b).
Interestingly, the length of many excitable
cells is of similar order. For example, in Loligo Pealei
squids the length of a giant axon is 3–5 times the length
of an AP (the axonal length is 10–20 cm, and the esti-
mated length of an AP is ≈4 cm (the pulse duration is
≈2 ms and the propagation velocity is ≈20 m/s) [11,54]).
Because the characteristics of APs over distances of the
order of the pulse size were scarcely investigated, it may
be worth to examine these properties in a future work.
Voltage signal of acoustic pulses. We now turn to
examine the effect of density changes on a transmem-
brane voltage measurement. Changes in electric prop-
erties, such as capacitance and resistance, are unavoid-
able during propagation of acoustic pulses in soft ma-
terials [31, 55]. Furthermore, the surface potential of a
lipid monolayer differs by ∼100 mV between the liquid-
expanded and the liquid-condensed phases, the same or-
der of magnitude as the electric aspect of an AP [24]. To
quantify these changes, we explore a simple extension to
the model; a standard electrophysiological measurement
across a lipid interface. For the moment we ignore any
transmembrane current of ions, and focus on currents
following solely from voltage changes generated by the
layer itself. The conservation of charge requires
(CV ) = 0,
d
dt
(8)
with C the local capacitance of the material, and V the
local transmembrane voltage. The equation implies an
inverse relation between the voltage and the material
capacitance
=
(9)
with Vi, Ci the initial voltage and capacitance, respec-
tively. The capacitance of the system can be associated
with the area, thickness and relative permittivity of the
material (A, d and ε, respectively)
V
Vi
Ci
C ,
C ≈ ε0ε
A
d
,
(10)
where all variables (including ε) depend on the local
thermodynamic state of the membrane [55]. Replacing
the area term with the density of the membrane, we ob-
tain a linear relation between the voltage and the local
density
C ∝ ε
dρ
=⇒ V ∝ dρ
ε
.
(11)
Thus, a propagating density pulse (Fig. 1), if detected
by a voltage sensor, should display a distinct shape that
6
Saturation(a)(b)(c)(d)Figure 5: Collision between two pulses as appeared in the (a) density, (b) pressure and (c) temperature fields. x- and y-axis represent
space and time, respectively. A local description of the (d) density, (e) pressure and (f) temperature is plotted as a function of time at
the collision point (x/L=0, solid line) and at some distance away from it (x/L=0.4, dashed-dotted line). The location of these solutions
in space is marked in (c) by solid and dashed-dotted line, respectively. Fluid initial density was ρ0=0.77. Excitation was conducted at
x0=±0.4π. Numerical calculation was performed with 8192 grid points, and the x-domain was [−0.4π, 0.4π]. Other parameters are similar
to those given in caption of Fig. 2.
includes a depolarization, repolarization and hyperpo-
larization phase, similar to measurements of APs. How-
ever, a voltage pulse is expected to display a sharper
non-linear response (a threshold) as compared with the
density pulse (Fig. 4b). This is because the relative per-
mittivity and thickness are also state dependent, the for-
mer decreases and the latter increases as the pulse tra-
verses into the condensed phase [24, 56, 57]. On top of
these results, one could include the resistance of the soft
system to transmembrane ionic currents, which is gov-
erned by a state dependent permeability [18,58]. In con-
clusion, the electric potential difference across a lipid in-
terface responds non-linearly when acoustic pulses travel
along the interface.
The realization that APs may be acoustic pulses,
that carry electrical changes as they travel, should have
significant implications for the field of neuroscience. For
decades the field has placed much focus on measure-
ments and analysis of electric signals in nervous-systems
of many organisms. However, it is possible that cru-
cial computational information goes unnoticed when the
state of a neuron is simplified into a raster plot data (a
digital-like uniformity of zeros and ones that represent
events of APs). Let us describe one plausible scenario.
In Fig. 4 we have demonstrated that at large excitation
amplitudes very similar density pulses can be excited
with a significantly different pressure signature. Thus, a
voltage sensor, sensitive to density but not to pressure,
would not resolve between different pressure signatures
that could induce different cellular responses, and re-
sult in a considerably different computational scheme as
compared to the classical electrical picture.
Model extensions. The evidence that the reported
pulses do not emerge from a fine-tuned model, but rather
result from an idealized description of a soft system, em-
phasizes the generality of the phenomenon. This, how-
ever, should only be viewed as a first step. To treat
pulses in soft systems more accurately, further exten-
sions to the model are required.
(1) The parameters
of the system; specifically, viscosity, heat capacity and
thermal conductivity, are state dependent and not con-
stants [19, 59, 60]. This is particularly evident near a
phase transition and should modify the detailed proper-
ties of the nonlinear pulses. (2) The van der Waals con-
stitutive equation is not an accurate representation of
the state diagram of soft materials, lipids in particular.
For example, in DPPC the volume exclusion appears at
≈0.8wc [24, 29, 40]. This is quantitatively different from
the van der Waals equation, where the volume exclu-
sion occurs at wc/3. These differences clearly influence
the properties of the pulse. For example, the saturation
7
(a)(b)(c)(d)(e)(f)amplitude of the density pulse would be ρsat ≈1.2ρc in-
stead of ≈3ρc. (3) The viscoelastic properties of lipids
are more complicated than simply considering a capillar-
ity term [48]. (4) No effect of geometry was considered
in this work (boundary conditions, extension to two- or
three-dimensions, coupling to bulk). (5) The parameters
that describe the material (critical point, thermal con-
ductivity, etc.) are very likely influenced by additional
components that were not considered here. For example,
pH, ions, as well as embedded proteins. (6) In order to
provide a detailed analysis of electrophysiological mea-
surements, a more accurate description of the electro-
mechanical coupling is necessary. For example, we did
not address space- and voltage-clamp experiments in this
paper [11]. (7) Our description is based on a mean field
approximation. Therefore, thermodynamic fluctuations
were not accounted. In particular, the model does not
describe current fluctuations that were experimentally
measured in non-living soft systems as well as in living
cells [18]. To accommodate for state fluctuations, an
analysis of the thermodynamic susceptibilities is neces-
sary. These aspects should be considered in a future
work.
5 Acknowledgements
The authors thank Christian Fillafer, Kevin Kang, Ju-
lian Kappler and Konrad Kaufmann for fruitful discus-
sions and valuable comments, and to Daniel Lecoanet for
support with the Dedalus open-source code. MM thanks
Uri Nevo for introducing him to the subject of non-
electric aspects of action potentials, and to Jay Fineberg
for useful comments. MM further acknowledges funds
from SHENC-research unit FOR 1543.
References
[1] I. Tasaki, Physiology and electrochemistry of nerve fibers.
New York: Academic press, 1982.
[14] A. T. Przybylski, W. P. Stratten, R. M. Syren, and S. W. Fox,
"Membrane, action, and oscillatory potentials in simulated
protocells," Naturwissenschaften, vol. 69, no. 12, pp. 561–
563, 1982.
[2] G. O. Mackie, "Epithelial conduction: Recent findings, old
questions, and where do we go from here?," Hydrobiologia,
vol. 530-531, pp. 73–80, 2004.
[15] I. Tasaki, "Rapid Structural Changes in Nerve Fibers and
Cells Associated With Their Excitation Processes," Japanese
Journal of Physiology, vol. 49, pp. 125–138, 1999.
[3] W. J. Parak, J. Domke, M. George, a. Kardinal, M. Rad-
macher, H. E. Gaub, a. D. de Roos, a. P. Theuvenet, G. Wie-
gand, E. Sackmann, and J. C. Behrends, "Electrically ex-
citable normal rat kidney fibroblasts: A new model system
for cell-semiconductor hybrids.," Biophysical journal, vol. 76,
no. 3, pp. 1659–1667, 1999.
[4] R. D. Fields, "Oligodendrocytes changing the rules: action
potentials in glia and oligodendrocytes controlling action po-
tentials," The Neuroscientist, vol. 14, no. 6, pp. 540–543,
2008.
[5] F. M. Ashcroft and P. Rorsman, "Electrophysiology of the
pancreatic β-cell," Progress in Biophysics and Molecular Bi-
ology, vol. 54, no. 2, pp. 87–143, 1989.
[6] S. P. Leys, "Elements of a 'nervous system' in sponges.," The
Journal of experimental biology, vol. 218, no. Pt 4, pp. 581–
91, 2015.
[7] M. J. Beilby, "Action potential in charophytes," International
review of cytology, vol. 257, pp. 43–82, jan 2007.
[8] C. L. Slayman, W. S. Long, and D. Gradmann, "?Action po-
tentials? in Neurospora crassa, a mycelial fungus," Biochim-
ica et Biophysica Acta (BBA)-Biomembranes, vol. 426, no. 4,
pp. 732–744, 1976.
[9] D. C. Wood, "Membrane permeabilities determining resting,
action and mechanoreceptor potentials in Stentor coeruleus,"
Journal of Comparative Physiology, vol. 146, no. 4, pp. 537–
550, 1982.
[10] J. M. Kralj, D. R. Hochbaum, A. D. Douglass, and A. E.
Cohen, "Electrical spiking in Escherichia coli probed with
a fluorescent voltage-indicating protein," Science, vol. 333,
no. 6040, pp. 345–348, 2011.
[11] A. L. Hodgkin and A. F. Huxley, "A Quantitative Descrip-
tion of Membrane Current and Its Application to Conduction
and Excitation in Nerve," Journal of Physiology, vol. 117,
pp. 500–544, 1952.
[12] D. J. Aidley and D. J. Ashley, The physiology of excitable
cells, vol. 4. Cambridge University Press Cambridge, 1998.
[13] S. Shrivastava and M. F. Schneider, "Evidence for 2D Soli-
tary Sound Waves in a Lipid Controlled Interface and its
Biological Implications for Biological Signaling," Journal of
The Royal Society Interface, vol. 11, pp. 1–8, 2014.
[16] I. Tasaki, A. Watanabe, and I. Singer, "Excitability of Squid
Giant Axon in the Absence of Univalent Cations in the Exter-
nal Medium," Proceedings of the National Academy of Sci-
ences, vol. 56, pp. 1116–1122, 1966.
[17] S. Terakawa, "Periodic responses in squid axon membrane
exposed intracellularly and extracellularly to solutions con-
taining a single species of salt," The Journal of Membrane
Biology, vol. 63, no. 1-2, pp. 51–59, 1981.
[18] B. Wunderlich, C. Leirer, A. L. Idzko, U. F. Keyser, A. Wix-
forth, V. M. Myles, T. Heimburg, and M. F. Schneider,
"Phase-state dependent current fluctuations in pure lipid
membranes," Biophysical Journal, vol. 96, no. 11, pp. 4592–
4597, 2009.
[19] T. Heimburg and A. D. Jackson, "On Soliton Propagation
in Biomembranes and Nerves," Proceedings of the National
Academy of Sciences, vol. 102, no. 28, pp. 9790–9795, 2005.
[20] K. Kaufmann, "Action Potentials and Electromechanical
Coupling in the Macroscopic Chiral Phospholipid Bilayer,"
1989.
[21] G. N. Ling, "A Revolution in the Physiology of the Living
Cell," 1992.
[22] A. L. Hodgkin and A. F. Huxley, "Resting and Action Poten-
tials in Single Nerve Fibers," J. Physiol., vol. 104, pp. 176–
195, 1945.
[23] J. P. Wikswo, J. P. Barach, and J. A. Freeman, "Magnetic
field of a nerve impulse: first measurements.," Science (New
York, N.Y.), vol. 208, no. 4439, pp. 53–55, 1980.
[24] D. Steppich, J. Griesbauer, T. Frommelt, W. Appelt, A. Wix-
forth, and M. F. Schneider, "Thermomechanic-electrical cou-
pling in phospholipid monolayers near the critical point.,"
Physical Review E, vol. 81, no. 6 Pt 1, p. 061123, 2010.
[25] J. Griesbauer, S. Bossinger, A. Wixforth, and M. F. Schnei-
der, "Propagation of 2D Pressure Pulses in Lipid Monolayers
and Its Possible Implications for Biology," Physical Review
Letters, vol. 108, no. 19, p. 198103, 2012.
[26] S. Shrivastava, K. H. Kang, and M. F. Schneider, "Solitary
shock waves and adiabatic phase transition in lipid interfaces
and nerves," Physical Review E, vol. 91, no. 1, pp. 1–7, 2015.
[27] B. Fichtl, S. Shrivastava, and M. Schneider, "Protons at the
speed of sound: Predicting specific biological signaling from
physics," Scientific Reports, vol. 6, no. 1, p. 22874, 2016.
8
[28] J. M. Ritchie and R. D. Keynes, "The production and ab-
sorption of heat associated with electrical activity in nerve
and electric organ," Quarterly Reviews of Biophysics, vol. 18,
no. 4, pp. 451–476, 1985.
[29] T. Heimburg, Thermal Biophysics of Membranes. Wiley-
VCH, 1st ed., 2007.
[30] J. Griesbauer, A. Wixforth, and M. F. Schneider, "Wave
propagation in lipid monolayers," Biophysical Journal,
vol. 97, no. 10, pp. 2710–2716, 2009.
[31] J. Griesbauer, S. Bossinger, A. Wixforth, and M. F. Schnei-
der, "Simultaneously Propagating Voltage and Pressure
Pulses in Lipid Monolayers of Pork Brain and Synthetic
Lipids," Physical Review E, vol. 86, no. 6, pp. 1–5, 2012.
[32] A. L. Hodgkin, A. F. Huxley, and B. Katz, "Measurement of
current-voltage relations in the membrane of the giant axon of
Loligo," The Journal of Physiology, vol. 116, no. 4, pp. 424–
448, 1952.
[33] I. Tasaki, "Collision of two nerve impulses in the nerve fibre,"
Biochimica et Biophysica Acta, vol. 3, pp. 494–497, 1949.
[34] R. Follmann, E. Rosa, and W. Stein, "Dynamics of sig-
nal propagation and collision in axons," Physical Review E,
vol. 92, no. 3, pp. 1–11, 2015.
[35] S. Shrivastava, K. H. Kang, and M. F. Schneider, "Collision
and Annihilation of Nonlinear Pulses and Action Potentials
in Interfaces," unpublished, 2017.
[36] J. Kappler, S. Shrivastava, M. F. Schneider, and R. R. Netz,
"Nonlinear fractional waves at elastic interfaces," arXiv:
1702.08864, 2017.
[37] M. Slemrod, "Dynamic phase transitions in a van der Waals
fluid," Journal of differential equations, vol. 52, no. 1, pp. 1–
23, 1984.
[38] M. Grinfeld, "Nonisothermal dynamic phase transitions,"
Quarterly of Applied Mathematics, vol. 47, no. 1, pp. 71–84,
1989.
[39] B. U. Felderhof, "Dynamics of the diffuse gas-liquid interface
near the critical point," Physica, vol. 48, no. 4, pp. 541–560,
1970.
[40] O. Albrecht, H. Gruler, and E. Sackmann, "Polymorphism
of phospholipid monolayers," Journal de Physique, vol. 39,
no. 3, pp. 301--313, 1978.
[41] R. Courant and K. O. Friedrichs, Supersonic flow and shock
waves, vol. 21. Wiley-Interscience, New York, 1948.
[42] A. N. Gorban and I. V. Karlin, "Beyond Navier?Stokes
equations: capillarity of ideal gas," Contemporary Physics,
vol. 7514, no. November, pp. 1–21, 2016.
[43] J. D. van der Waals, "The thermodynamic theory of capil-
larity flow under the hypothesis of a continuous variation of
density (Verhandel/Konink. Akad. Weten., 1893, vol. 1, En-
glish Translation)," Journal of Statistical Physics, vol. 20.
[44] L. D. Landau and E. M. Lifshitz, "Statistical physics, vol. 5,"
Course of theoretical physics, vol. 30, 1980.
[45] J. W. Cahn and J. E. Hilliard, "Free energy of a nonuniform
system. I. Interfacial free energy," The Journal of chemical
physics, vol. 28, no. 2, pp. 258–267, 1958.
[46] D. J. Korteweg, "Sur la forme que prennent les ´equations du
mouvement des fluides si l'on tient compte des forces capil-
laires caus´ees par des variations de densit´e consid´erables mais
continues et sur la th´eorie de la capillarit´e dans l'hypothese
d'une variatio," Archives N´eerlandaises des Sciences exactes
et naturelles, vol. 6, no. 1, p. 6, 1901.
[47] D. C. Johnston, "Thermodynamic properties of the van der
Waals fluid," arXiv:1402.1205, 2014.
[48] G. Espinosa, I. L´opez-Montero, F. Monroy, and D. Langevin,
"Shear rheology of lipid monolayers and insights on mem-
brane fluidity," Proceedings of the National Academy of Sci-
ences, vol. 108, no. 15, pp. 6008–6013, 2011.
[49] J. S. Clegg and W. Drost-Hansen, "On the biochemistry and
cell physiology of water-Chapter 1," 1991.
[50] D. J. Benvegnu and H. M. Mcconnell, "Line Tension between
Lipid Domains in Lipid Monolayers," Journal of Physical
Chemistry, vol. 96, pp. 6820–6824, 1992.
[51] K. J. Burns, G. M. Vasil, J. S. Oishi, D. Lecoanet, B. P.
Brown, and E. Quataert, "Dedalus: A Flexible Framework
for Spectrally Solving Differential Equations," (unpublished),
2017.
[52] D. J. Bakkum, U. Frey, M. Radivojevic, T. L. Russell,
J. Muller, M. Fiscella, H. Takahashi, and A. Hierlemann,
"Tracking axonal action potential propagation on a high-
density microelectrode array across hundreds of sites.," Na-
ture communications, vol. 4, p. 2181, 2013.
[53] F. Patolsky, B. P. Timko, G. Yu, Y. Fang, A. B. Greytak,
G. Zheng, and C. M. Lieber, "Detection, stimulation, and in-
hibition of neuronal signals with high-density nanowire tran-
sistor arrays," Science, vol. 313, no. 5790, pp. 1100–1104,
2006.
[54] J. Z. Young, "The structure of nerve fibres in Cephalopods
and Crustacea," Proceedings of the Royal Society B: Biolog-
ical Sciences, vol. 121, no. 823, pp. 319–337, 1936.
[55] T. Heimburg, "The capacitance and electromechanical cou-
pling of lipid membranes close to transitions: The effect
of electrostriction," Biophysical Journal, vol. 103, no. 5,
pp. 918–929, 2012.
[56] T. Heimburg, "Mechanical aspects of membrane thermo-
dynamics. Estimation of the mechanical properties of lipid
membranes close to the chain melting transition from
calorimetry," Biochimica et Biophysica Acta - Biomem-
branes, vol. 1415, no. 1, pp. 147–162, 1998.
[57] Y. Kimura and a. Ikegami, "Local dielectric properties
around polar region of lipid bilayer membranes.," The Jour-
nal of membrane biology, vol. 85, no. 3, pp. 225–231, 1985.
[58] E. El-Mashak and T. Tsong, "Ion selectivity of temperature
- induced and electric field induced pore in dipalmitoylphos-
phatidylcholine vesicles," Biochemistry, vol. 24, pp. 2884–
2888, 1985.
[59] E. Hermans and J. Vermant, "Interfacial shear rheology of
DPPC under physiologically relevant conditions.," Soft mat-
ter, vol. 10, pp. 175–186, 2014.
[60] S. Youssefian, N. Rahbar, C. R. Lambert, and S. V. Des-
sel, "Variation of thermal conductivity of DPPC lipid bilayer
membranes around the phase transition temperature," J. R.
Soc. Interface, vol. 14, p. 20170127, 2017.
Supplemental materials
A Model Equations
The supplemental materials are intended to provide in-
terested readers with further mathematical details (sec-
tion A), additional solutions (section B) and caption to
the supplemental movie (section C).
9
Model variables and parameters. The fluid con-
sidered in this letter is described by six dynamic vari-
ables of space and time, that are coupled to one another
by three conservation laws (Eq. (1)), two constitutive
equations (Eq. (4)), and one inverse relation, ρ = w−1.
The six variables are listed in Table S1.
In addition,
the model depends on seven constant parameters, all
are measurable physical quantities (Table S2). The van
Variable
Name
Units for a 3d material
(2d material) in MKS
system
w
v
E
p
θ
ρ
specific volume
m3
kg
velocity
m
s
specific total energy m2
s2
pressure
temperature
density
N
m2
K
kg
m3
(cid:17)
(cid:17)
kg
(cid:16) m2
(cid:0) N
(cid:1)
(cid:16) kg
m
m2
Table S1: Model variables
Parameter
Name
Units for a 3d material
(2d material) in MKS
system
m
a
b
ζ
cv
k
C
kg
mass of a fluid particle
volume exclusion by the particles
average attraction between fluid particles N m4(cid:0)N m3(cid:1)
m3(cid:0)m2(cid:1)
(cid:17)
(cid:16) kg
(cid:0) J
(cid:16) kg2
specific heat capacity at constant volume
thermal conductivity
bulk viscosity
(cid:1)
(cid:17)
s
capillarity coefficient
kg
m·s
J
kg·K
J
m·s·K
kg2
m2·s2
s·K
s2
Table S2: Model constant parameters
10
Figure S1: Comparison of pulse shape at two different values of the capillarity coefficient, C/ζ2 = 10−20 (solid line) and 1 (dotted-dashed
line). Excitation amplitude was pexc = 60 and 150, respectively. Duration of excitation was t0 = 0.3 for both curves. Pulse was measured
at x/L=0.5 and 0.8, respectively. Other parameters appear in caption of Fig. 2.
der Waals parameters (m, a, b) are related to the critical
point according to
Note that a physically meaningful solution requires
1/3 < w and 0 < θ.
wc =
3b
m
,
pc =
a
27b2 ,
θc =
8a
27bkB
.
(S1)
The scaling of the spatial coordinate h (in the La-
grange frame) was defined according to the critical den-
sity ( ¯w = wc). Finally, kB
Boltzmann constant.
∼= 1.38 · 10−23 J/K is the
Model equations in the Euler frame. The model
equations in the Lagrange frame are given in Eq. (1).
These equations were transformed into the Euler frame
using Eq. (2); i.e., ∂x/∂h = w/ ¯w, and by replacing the
time derivative with the material derivative ∂t → Dt =
∂t + v∂x
Dimensionless model equations. The critical point
and the fluid viscosity parameters were used (Eq. (5))
to introduce the following reduced variables and param-
eters:
X ≡ h
L
,
t ≡ t
T
w ≡ w
wc
θ ≡ θ
θc
,
,
v ≡ v
U
p ≡ p
pc
,
,
E ≡ E
U 2 ,
τ ≡ τ
pc
.
(S2)
and
The spatial dimension in the Lagrange frame was nor-
malized by the choice ¯w = wc. The dimensionless model
equations are
Dtw = w∂xv,
Dtv = w∂x(τ1 + τ2),
(cid:20)
DtE = w
∂x(τ1v) +
(cid:21)
,
k
¯w2 w∂x(w∂xθ)
w
¯w
τ1 = −p + ζ
τ2 = −C
w
¯w2 ∂x(w∂xw).
∂xv,
(S7)
(S8)
In comparison, the classical Navier-Stokes equations in
the Euler frame are
with
and
∂t w = ∂X v,
∂tv = ∂X (τ1 + τ2) ,
E = ∂X (τ1v) + k∂2
∂t
X
τ1 = −p + ∂X v,
τ2 = − C∂2
X w,
θ,
(S3)
(S4)
with
Dtw = w∂xv,
Dtv = w∂xτN S,
DtE = w[∂x(τN Sv) + k∂2
xθ],
τN S = −p + ζ∂xv.
(S9)
(S10)
p =
E = cv
8θ
3 w − 1
θ − 3
w
− 3
w2 ,
v2
2
+
.
(S5)
B Pulse solution in a vdW fluid
model
The model equations (S3)–(S5) depend on three dimen-
sionless parameters: heat capacity, thermal conductivity
and capillarity coefficient
cv =
cvθc
pcwc
,
k =
kθc
pcwcζ
,
C =
C
ζ 2 .
(S6)
Capillarity coefficient and the line tension.
Gibbs approached a phase boundary as a infinitesimal
line that separates the two phases. The energy stored
in a phase boundary of length L and line tension γ is
E = γL. In contrast, van der Waals approached a phase
11
Figure S2: (a) Variation of stimulation duration (t0) did not affect the duration of the compression part of the density pulse, as measured
at distance x/L=1 from the excitation point. The subsequent rarefaction, however, was altered. (b) Density, (c) pressure and (d) temper-
ature solutions as a function of space (x-axis) and time (y-axis) upon stimulation with t0/T=0.4. Pulse maintains a recognizable shape
for distances larger than 10L. (e) The shape of the pulse did not change much with an increase in stimulation width. Initial density was
ρ0 = 0.67, and numerical calculation was conducted with 2048 grid points, x-domain [−5π, 5π] and dt=10−3. Other parameters appear in
the caption of Fig. 2.
transition as a continuous change, and the phase bound-
ary, therefore, has a finite length. Van der Waals sug-
gested an energy term that depends on the density gra-
dient
E =
K
d2x,
(S11)
(cid:90)
(cid:18) ∂ρ
(cid:19)2
∂x
with K ∼ C/ρ3
c. For a phase boundary of width (cid:96), the
energy stored in the phase boundary is E ∼ K ∆ρ2
(cid:96)2 (cid:96)L.
Therefore, the capillarity coefficient is related to the line
tension according to
C ∼ γ(cid:96)ρ3
∆ρ2 .
c
(S12)
Assuming (cid:96) ∼ 10−9 − 10−6 m, ∆ρ ∼ ρc ∼ 10−6 kg/m2
and γ ∼ 10−12 N (Benvegnu and McConnell, J Phys
Chem, 1992) we obtain C ∼ 10−27−10−24 kg2/s2. How-
ever, in the numerical calculation we have used a larger
value (C/ζ 2 ∼ 1), to avoid sharp gradients in density.
This did not have qualitative effect on the properties of
sound pulses near phase transition as evident in Fig. S1.
Effect of stimulation parameters on the pulse
characteristics. Figure S2a shows the effect of stim-
12
ulation duration on the shape of the density pulse, as
measured at distance x/L=1.
It was evident that the
compression part of the density pulse was not affected
much and maintained a pulse duration of ≈0.1T. On
the other hand, the subsequent rarefaction monoton-
ically decreased its density with stimulation duration.
The density, pressure and temperature field solutions as
a function of space (x-axis) and time (y-axis) are shown
in Figs S2b,c,d, respectively, for stimulation duration of
t0/T=0.4. It was evident that the pulse maintained a
recognizable shape for distances larger than 10L. An in-
crease of the stimulation width (λ) resulted in similar
observations (Fig. S2e).
C Caption
of
supplemental
movie
Movie S1: Collision between two propagating pulses is
demonstrated in the density, pressure and temperature
fields. At the peak of the collision a short burst of pres-
sure is obtained (t/T=0.2). The pulses annihilate af-
terwards. Computational details are given in caption of
Fig. 5.
t /T=0.40(c)(d)(e) |
1702.00350 | 1 | 1702 | 2017-01-30T16:12:55 | Effect of iron oxide loading on magnetoferritin structure in solution as revealed by SAXS and SANS | [
"physics.bio-ph",
"cond-mat.soft"
] | Synthetic biological macromolecule of magnetoferritin containing an iron oxide core inside a protein shell (apoferritin) is prepared with different content of iron. Its structure in aqueous solution is analyzed by small-angle synchrotron X-ray (SAXS) and neutron (SANS) scattering. The loading factor (LF) defined as the average number of iron atoms per protein is varied up to LF=800. With an increase of the LF, the scattering curves exhibit a relative increase in the total scattered intensity, a partial smearing and a shift of the match point in the SANS contrast variation data. The analysis shows an increase in the polydispersity of the proteins and a corresponding effective increase in the relative content of magnetic material against the protein moiety of the shell with the LF growth. At LFs above ~150, the apoferritin shell undergoes structural changes, which is strongly indicative of the fact that the shell stability is affected by iron oxide presence. | physics.bio-ph | physics | Colloids and Surfaces B: Biointerfaces 123 (2014) 82–88
Effect of iron oxide loading on magnetoferritin structure in solution as
revealed by SAXS and SANS
L. Melníková1, V.I. Petrenko2,3, M.V. Avdeev2, V.M. Garamus4, L. Almásy5, O.I. Ivankov2,3,
L.A. Bulavin3, Z. Mitróová1, P. Kopčanský1
1Institute of Experimental Physics, SAS, Watsonova 47, 040 01 Kosice, Slovakia
2Joint Institute for Nuclear Research, Joliot-Curie 6, 141980 Dubna, Moscow region, Russia
3Kyiv Taras Shevchenko National University, Volodymyrska Street 64, Kyiv, 01033 Ukraine
4Helmholtz-Zentrum Geesthacht: Centre for Materials and Coastal Research, Max-Planck-Street
5Wigner Research Centre for Physics, HAS, H-1525 Budapest, POB 49, Hungary
1, 21502 Geesthacht, Germany
Corresponding author: [email protected], tel.: +421 55 792 2233, Fax: +421 55 633 62 92
KEYWORDS: magnetoferritin, apoferritin, biorelevant magnetic nanoparticles, small-angle
neutron scattering, small-angle X-ray scattering, contrast variation
ABSTRACT
Synthetic biological macromolecule of magnetoferritin containing an iron oxide core inside a
protein shell (apoferritin) is prepared with different content of iron. Its structure in aqueous
solution is analyzed by small-angle synchrotron X-ray (SAXS) and neutron (SANS) scattering.
The loading factor (LF) defined as the average number of iron atoms per protein is varied up to
LF=800. With an increase of the LF, the scattering curves exhibit a relative increase in the total
scattered intensity, a partial smearing and a shift of the match point in the SANS contrast
variation data. The analysis shows an increase in the polydispersity of the proteins and a
corresponding effective increase in the relative content of magnetic material against the protein
moiety of the shell with the LF growth. At LFs above ~150, the apoferritin shell undergoes
structural changes, which is strongly indicative of the fact that the shell stability is affected by
iron oxide presence.
1
1. INTRODUCTION
Apoferritin being a part of the natural biological macromolecule of ferritin [1] represents
a very useful confinement of magnetic nanoparticles inside for biomedical applications [2,3].
This almost spherical protein shell with an external diameter of 12 nm and thickness of about
2.5 nm makes it possible to disperse nanoparticles (by placing them in its cavity) in biological
media and additionally minimize their possible toxic effect. It also prevents the bulk aggregation
of nanoparticles and restricts their maximal size. In case of magnetic nanoparticles (Fe3O4, γ-
Fe2O3) placed inside an apoferritin shell the corresponding protein is known as magnetoferritin
[4]. It is of current interest for various biomedical applications, which make use of the magnetic
properties of nanoparticles, such as targeted drug transport, magnetic resonance imaging, etc.
[5,6]. In addition to biocompatibility another advantage of magnetoferritin is the relatively short
time of synthesis, in which the magnetite (Fe3O4) is formed inside the protein cavity. Through
the regulation of the iron-to-apoferritin ratio it is possible to prepare homogeneously dispersed
magnetoferritin molecules with different iron oxide loading. Number of iron atoms per one
molecule of protein shell is referred to as a loading factor (LF) [7].
In the previous studies the structure characterization of the magnetic core at various LFs
was performed mostly by transmission electron microscopy (TEM). In particular, an increase in
the magnetic nanoparticle size with the LF growth and non-spherical core shapes for low LFs
were reported [7,8]. It was also found that for quite high LFs (>1000) particle aggregates are
formed [7,9]. By separating and extracting the non-aggregated particles having magnetic core,
the uniform magnetoferritine molecules can be crystallized to a 3D ordered magnetic array [10].
A number of experiments have been done to characterize the mineral composition of a synthetic
magnetoferritin core. Thus, Mössbauer spectroscopy showed [11] that it is rather different from
that of native ferritin. Faraday rotation measurements showed [12] that the composition of the
core changes with increasing LF starting from maghemite with a relatively small fraction (about
10%) of magnetite at LF <1250 and varying towards 100 % of magnetite at LFs > 3250.
One can see that the previous studies of magnetoferrtin were mainly focused on the
samples with LFs approaching the upper limit of the possible iron oxide content within the
protein shell. However, recent investigations showed [13,14] that some structural changes of the
protein shell, and also the organization of magnetoferritin in solutions are observed already at
significantly lower LFs. The present paper aims at studying the influence of the magnetic content
of magnetoferritin on the structure of the protein shell at low and moderate iron oxide loadings
2
(LF < 800) by small-angle X-rays (SAXS) and neutron (SANS) scattering in order to provide
additional and detailed characterization of this novel material and to follow the possible
modifications of the protein cage in a wide interval of the iron oxide loading. Both kinds of the
small-angle scattering technique cover the length scale of 1-100 nm, but have different
sensitivity to the same elements. Especially, it concerns hydrogen whose replacement with
deuterium provides wide possibilities of the so-called contrast variation in SANS. SAXS, in turn,
is highly sensitive to the heavier atoms, such as iron. Here, the general size characteristics of
magnetoferritin and their aggregates in aqueous solutions are first obtained by analyzing SAXS
and SANS data complemented by the dynamic light scattering (DLS) measurements. Then, the
SANS contrast variation is additionally applied based on the mixtures of heavy and light water to
conclude about the composition of magnetoferritin.
2. MATERIALS AND METHODS
Natural apoferritin (horse spleen) was purchased from Sigma-Aldrich. Magnetoferritin
with various LFs up to 800 was synthesized in anaerobic conditions at 65°C and alkaline pH as
described in details elsewhere [15,16]. First, apoferritin was added into the AMPSO (3-[(1,1-
Dimethyl-2-hydroxyethyl)amino]-2-hydroxy-propanesulfonic acid) buffer (0.05 M AMPSO
buffered with 2 M NaOH to pH 8.6) to achieve protein concentration 6 mg∙mL-1. The buffer was
deaerated for 55 minutes with nitrogen and for further 5 minutes after the addition of apoferritin.
Then the solution in the reaction bottle was hermetically closed and placed in preheated water
bath with temperature 65°C on a magnetic stirrer. Next, deaerated solutions of reactants,
trimethylamine N-oxide and ferrous ammonium sulfate hexahydrate, were added dropwise into
the reaction bottle. After the synthesis all samples were filtered through 200 nm filter to remove
possible aggregates. The average loading factor of each sample was determined using UV-VIS
spectrophotometer SPECORD 40 (Analytik Jena, Germany). Protein concentration was
determined using the standard Bradford method at wavelength (λ) 595 nm and the amount of
iron was measured after HCl/H2O2 oxidation and KSCN addition by light absorption of
thiocyanate complex at λ = 400 nm.
Dynamic light scattering measurements were made on a Zetasizer Nano ZS 3600
(Malvern Instruments) at 25°C. The samples were diluted with 0.15 M NaCl to achieve the
protein concentration of ~ 0.2 mg∙mL-1 and filtered through a 0.2 µm syringe filter before
measurement.
For SANS contrast variation experiments magnetoferritin samples were freeze-dried for
24 hours after the synthesis to obtain a powder. 10 mg∙mL-1 solutions regarding the protein
3
concentration were prepared by dissolving powders in H2O/D2O mixtures with varying the D2O
volume fraction. The mixtures of AMPSO buffer (0.05 M AMPSO buffered with 2 M NaOH to
get pH 8.6) with the same ratios of H2O/D2O as in the samples, were used as background
solutions.
SAXS experiments were performed at the P12 BioSAXS beamline of the European
Molecular Biology Laboratory (EMBL) at the storage ring PETRA III of the Deutsche
Elektronen Synchrotron (DESY, Hamburg, Germany) at 20°C using a Pilatus 2M detector
(14751679 pixels) (Dectris, Switzerland) and synchrotron radiation with a wavelength
λ = 0.1 nm. The sample-detector distance was 3 m, allowing for measurements in a q-range of
0.11-4.4 nm-1. The q-range was calibrated using the diffraction patterns of silver behenate. The
experimental data were normalized to the transmitted beam intensity, corrected for a non-
homogeneous detector response, and the background scattering of the aqueous buffer was
subtracted. An automatic sample changer for a sample volume of 15 μL was used. The
experimental time including sample loading, exposure, cleaning and drying was about 1 min per
sample. The solvent scattering was measured before and after the sample scattering in order to
control the eventual sample holder contamination. Four consecutive frames comprising the
measurements for the solvent, the sample, and the solvent were taken. No measurable radiation
damage was detected by comparing four successive time frames with 5 s exposures. The final
scattering curve was obtained using the automated acquisition and analysis by averaging the
scattering data collected from different frames [17].
SANS measurements were carried out at the small-angle diffractometers SANS-II at the
SINQ spallation neutron source (PSI, Villigen, Switzerland) [18], operating in continuous
regime, and YuMO at the IBR-2 pulsed reactor (JINR, Dubna, Russia), in time-of-flight regime.
On SANS-II the scattering data were recorded at sample-detector distances of 1.3 and 4 m, with
a neutron wavelength of 0.53 nm and wavelength spread of about 10%. The raw data were
corrected for background, transmission and detector efficiency, and put on the absolute scale
using the scattering from a 1-mm thick H2O sample, pre-calibrated by scattering from a dilute
solution of polystyrene. The data were reduced by the BerSANS software package [19]. On the
YuMO small-angle spectrometer a two-detector set-up with ring wire detectors were used [20].
The neutron wavelength range was 0.05-0.8 nm-1. The measured scattering curves were corrected
for background scattering from buffer solutions. For absolute calibration of the scattered
intensity during the measurements a vanadium standard was used. The raw data treatment was
performed by the SAS program with a smoothing mode [21]. For the measurements on both
4
instruments the solutions were put in 1 mm thick quartz plain cells (Helma) and kept at room
temperature.
3. RESULTS AND DISCUSSIONS
3.1. SAXS and SANS data analysis at full contrast
As the first step, the SAXS scattering curves for apoferritin and magnetoferritin with the
minimal LF of 160 are compared in Fig. 1. The scattering curve of apoferritin (LF = 0) is well
described by the form-factor of a monodisperse spherical shell [22, 23]:
P(q) = (1/V)2 [V1(qR1) – V2(qR2)]2,
(1)
where (x) = 3(sin(x) – x cos(x)) / x3; Vi = (4/3)Ri
3 is the volume of a sphere with radius Ri; R1
= 6.32 (1) nm and R2 = 3.53 (1) nm are the outer and inner radii of the shell, respectively, and V
= V1 – V2 is the volume of the shell. The logical extension of this model for the case of
magnetoferritin is the representation of the macromolecule as a spherical particle with a
homogeneous, iron oxide containing core and the protein shell. However, this approach cannot
describe the experimental data obtained, most likely because of modification of the
magnetoferritin structure under iron oxide loading, which breaks the spherical symmetry.
Therefore, the data treatment is mainly reduced to the comparative analysis of specific
characteristics of the curves including the scattering invariants. Thus, in small-angle scattering at
sufficiently small q-values one can use the Guinier approximation:
(2)
where the forward scattered intensity
is determined by the particle number
density, n, particle volume, V, and the contrast,
which is the difference between the
mean scattering length densities (SLDs) of the particle,
, and solvent,
; and
is the radius
of gyration, the average of square distances from the center-of-mass of the macromolecule
weighted by the SLD distribution. Since apoferritin, like most of proteins, is a homogeneous
object in terms of the inner SLD fluctuations, its radius of gyration is strictly determined by the
inner and outer radii of the protein shell,
, which gives Rg =
5.25 nm well testified by the direct approximation of Eq.2 to the experimental curve in the
Gunier region (q < 0.3 nm-1).
The scattering curve of magnetoferritin solution retains its character typical for a spherical
shell, but an appreciable smearing of the peaks and a shift of the minima (indicated by arrows in
Fig. 1) towards larger q-values are observed. The radius of gyration of magnetoferritin found
5
22()(0)exp(/3)gIq=IRq22(0)(Δ)I=nVρΔsρ=ρρρsρgR255331212(3/5)()/()gRRRRRfrom the Guinier approximation to the experimental curve, 4.99 nm, is slightly smaller than that
of apoferritin. Also, the total intensity is larger than that for apoferritin with the same
concentration of protein moiety in the solution; in particular, the forward scattered intensity of
magnetoferritin exceeds that of apoferritin by 1.7 times. The observed differences cannot be
attributed to a simple transformation of the hollow apoferritin shell into a core-shell structure
after the cage is filled with iron oxide. First, the scattering curve of magnetoferritin cannot be
properly described in terms of a simple model of monodisperse core-shell spheres as such model
cannot principally explain the observed smearing. Second, the measured increase in the forward
scattered intensity of magnetoferritin is too high; the volume fraction of magnetic material in the
system at LF=160 is at the level of 0.005, which should give maximum a 10 % increase in the
squared contrast relative to apoferritin, which is much below the observation. Since the protein
shell is monodisperse, the discussed increase in the intensity suggests that the magnetic material
has a non-uniform distribution over the protein shells. It was reported previously [7,8] that the
loading of magnetoferritin similar to native ferritin [24,25] is characterized by some distribution
of the iron content over the cages. From the viewpoint of the scattering theory one deals in this
case with a distribution of
with the mean value,
(effective mean SLD), and width, p,
which determines the so-called structural polydispersity [25-27] and gives an additional
contribution to the scattering. In particular, for the forward scattered intensity one can write:
,
(3)
where the modified (for polydisperse systems) contrast is determined as
. Using the
experimentally found ratio between the forward scattered intensities for apoferritin and
magnetoferritin at LF=160 one obtains p = 0.07 e∙Å-3 (here for SLD in SAXS we use
traditionally the units of number of electrons per volume). This is more than ten times larger than
the difference in the mean SLD of magnetoferritin with LF=160 (
= 0.425 e∙Å-3) and
apoferritin (
= 0.42 e∙Å-3). Thus, the volume fraction of the magnetic material in the cage
varies in a much wider interval than what can be achieved for magnetoferritin regarded as a
monodisperse protein cavity with just varying amount of iron oxide. This contradiction suggests
that the protein shell is partly disassembled. The shell in this case is no longer a monodisperse
object, and now, in addition to the structural polydispersity, the size polydispersity contributes to
the scattering as well. The partial disassembling of the shell is indirectly confirmed by modelling
the scattering curves by indirect Fourier transform (IFT) [28], using the GNOM program [29],
which represents the scattering data in terms of the pair distance distribution function (PDD) (see
6
ρeρ2222(0)pI=nΔρV+nσVseρρ=ρΔ~ρρinset to Fig. 1). The PDD function for magnetoferritin differs significantly from that of
apoferritin, which is strongly skewed towards large distances owing to the protein shell around
the empty cavity. Still, the maximal sizes are close for the two macromolecules. From the
comparison of the PDD functions of magnetoferritin and a filled sphere with the diameter of
apoferritin (calculated and plotted additionally in inset to Fig. 1) one can conclude that the
scattering object in our case has an intermediate shape between spherical shell and sphere. This
conclusion is also supported by the ab-initio analysis of the scattering data using the DAMMIF
program [30], which models the shape of the scattering object in the homogeneous
approximation by representing it with a set of sufficiently small uniform beads (Fig. 2). As
compared to the scattering from apofferitin, for which DAMMIF, as expected, gives a shape very
close to a hollow sphere (Fig. 2a), the DAMMIF treatment of the scattering from magnetoferritin
results in a structure, which deviates strongly from a complete shell (Fig. 2b). It must be noted
that this structure is some kind of an average shape, which does not exclude the existence of
complete shells in the solution. The given treatment fully neglects the scattering contribution
from magnetite. Still, it demonstrates clearly that the explanation of the observed shifts in the
scattering minima and smearing of the curves requires quite significant deviations from a hollow
sphere.
The increase in LF is accompanied by further smearing of the SAXS curves, as one can
see in Fig. 3a which covers intermediate loading factors up to LF = 430. This is reflected in the
PDD functions obtained by the IFT procedure (Fig. 3b) as a shift of the particle peak to smaller
distances, which corresponds to a spherical symmetry violation and a transition to a more
compact object. Along with it, the character of the curves changes as well, showing some
specific increase in the forward scattered intensity and the radius of gyration both obtained as a
result of the IFT procedure (Fig. 4). The latter is an indication of the formation of aggregates of
magnetoferritin in the solutions with the LF growth. This is reflected in the corresponding PDD
functions (Fig. 3b) as the appearance of a wide band above r = 12 nm (the expected diameter of
the complete protein shell) starting from LF = 260. The ratios between the calculated and
measured values of I(0) and Rg correspond to rather small (< 10) aggregation numbers. Such
aggregation alone cannot explain the observed smearing of the curves, hence, it points to the
increasing polydispersity with increasing amount of magnetic material in magnetoferritin. The
size and structure polydispersity together with the absence of strictly defined scattering form-
factor of the macromolecules prevent the easy determination and separation of the structure-
factor which would correspond to the average effective interactions of the basic structural units
(here, magnetoferritin macromolecules) in the solutions like in the case of homogeneous or
7
multilayered structures [31, 32]. Instead, the structures formed resemble more the partly
aggregated particles in aqueous dispersions of magnetite nanoparticles coated with surfactant
shells [33].
The SANS curves (Fig. 5a) and the corresponding PDD functions (Fig. 5b) for another
series of magnetoferritin solutions cover a more extended interval over LFs, up to LF = 800. For
the intermediate LFs (LF < 600) the similar treatment generally repeats the previous conclusions
of the SAXS analysis; yet, the aggregate effect starts to be visible at higher LF and is
characterized by smaller aggregate size for the second series. At the same time, starting from LF
= 600 a tendency towards a sharp increase in the aggregation is seen, which is well distinguished
as a drastic widening of the corresponding PDD functions (Fig. 5b). A further increase in LF
would make it impossible to treat the curves in the same way at the given instrumental
resolution, which is determined by the minimal q-value corresponding to the detectable maximal
size of the scattering objects. The formation of stable aggregates in magnetoferritin solutions
with the LF growth is confirmed by the DLS measurements from diluted solutions. In Fig. 6 the
LF-dependences of the mean hydrodynamic radius, <Rhydr>, and of radius of gyration obtained
by SAXS and SANS are compared. For apoferritin the <Rhydr>-value is fully consistent with the
small-angle scattering data if one takes into account that in this case, the radius of gyration
corresponds to the radius of the hollow protein shell of about 6 nm, and the hydrodynamic radius
naturally exceeds this value by about 10%. A non-monotonic size growth is revealed in the three
kinds of experiments, as shown in Fig. 6 from which one can reliably conclude that a tendency to
a slight aggregation of magnetoferritin is seen at LF over the interval of 160 – 510, and the
aggregation becomes more intensive at LF above 600. The discussed LF-dependences do not
fully repeat themselves most probably because of a strong sensitivity to stochastic factors during
the sample preparation (e.g. intensity and time of solution deaeration and stirring), which is
typical for liquid dispersions of nanoparticles.
It is interesting to compare the size characteristics of magnetoferritin with those of ferritin
(Figs. 3, 5, 6) in solutions under the same conditions. The natural LF-values of ferritin (the core
has a ferrihydrite-like structure) are close to LF = 2000. One can see that despite the large iron
content the scattering curves from ferritin show more pronounced oscillating behaviour, thus
reflecting rather high monodispersity and structural stability of this macromolecule. At the same
time, the corresponding PDD functions (Figs. 3 b, 5 b) indicate that the ferritin solutions are not
free of some small aggregates; still, their mean size is significantly lower as compared to the
solutions of magnetoferritin.
8
The disassembling of the protein shell in apoferritin can take place under some
conditions, in particular in strongly acidic solutions [34]. Thus, it was shown that the complete
disappearance of the characteristic peaks in SAXS curves from disassembled apoferritin strictly
takes place when 12 out of the 24 structural units are removed from the shell. In our experiments
pH was kept constant at 8.6, which is optimal for the stability of apoferritin structure, but the
character of the observed smearing of the scattering curves was the same, thus indicating that in
average about half of the apoferritin shell in magnetoferritin is destroyed when LF approaches
1000.
3.2. SANS contrast variation
The contrast variation technique in SANS experiments on moderately polydisperse
objects makes it possible to conclude about the polydispersity degree in terms of the weighted
averaged scattering length density distribution over the studied particles [26]. For this purpose
the scattering from the system under study is analysed, varying the content of a deuterated
component of the solvent. Here, the SANS contrast variation data based on substitution of light
(H2O) for heavy (D2O) water is used to conclude about the change in the polydispersity with
rising LF. The samples with low (LF = 160) and relatively high (LF = 510) loading factors were
investigated. The upper LF-value was chosen to avoid the large aggregation which starts, as
shown above, at LF of about 600. The Guinier region for the different contrasts (Fig. 7) was used
to determine the I(0) parameter according to Eq. 1. I(0) as a function of the volume fraction of
D2O is shown in Fig. 8; its minimum for the case of polydisperse particles corresponds to the
effective match point [26]. At the considered LF-values the additional magnetic neutron
scattering contribution can be neglected. The upper estimates give for LF=510 that its
contribution to the forward scattered intensity is less than 2%. Already in the monodisperse
approximation, under the assumption that the magnetic core in magnetoferritin consists of
magnetite (Fe3O4, SLD = 6.9∙1010 cm-2), the shifts of the effective match points (the
corresponding SLDs are 2.46∙1010 cm-2 and 2.79∙1010 cm-2 for LF=160 and LF=510,
respectively) as compared to the protein moiety of apoferritin (SLD 2.34∙1010 cm-2) give 0.026
and 0.099, respectively, for the volume fractions of magnetic material in the protein cage. These
values are much larger than the amount of iron loaded during the synthesis (0.005 and 0.017 for
LF=160 and LF=510, respectively). Assuming maghemite (Fe2O3, SLD = 6.7∙1010 cm-2) in the
magnetic core, the result would differ by less than 5%.
The obtained match points are significantly higher than those expected for a core filled
with iron oxides. Therefore, the SANS contrast variation strongly points to an abnormally high
9
average ratio between the content of the magnetic material and protein, which can be explained
by the partial disassembling of the shell, leading to an effective growth of the relative content of
the magnetic component in the structure of magnetoferritin. The residual scattering in the
effective match points, which is an indicator of polydispersity, increases, thus explaining the
broadening of the polydispersity for larger LFs. This is consistent with the smearing of the
scattering curves with the LF growth in Figs. 3a, 5a.
It should be noted that despite the concluded disassembling of the protein shell the
magnetoferritins remain soluble as a whole. Also, the solutions themselves are stable for at least
three months, without signs of precipitates. The mechanism of the effect of magnetic loading on
the protein structure is unclear. As mentioned previously, apoferritin disassembly was observed
at pH below 3.4 [34], which, however, is not our case, since magnetoferritin was prepared in
alkaline pH 8.6 and anaerobic conditions. While it is not possible to control pH directly during
the synthesis process, after the synthesis the pH value was checked and only slight decrease for
higher LFs was detected, remaining above pH 7 for all LFs.
One can relate the observed structural change of the protein with a specific effect of
magnetic nanoparticles placed in the cavity on the protein shell. So far there is no general
understanding of interactions between nanoparticles and proteins despite of the extensive studies
of this problem in recent years. In particular, the interaction of various nanoparticles with
specific protein aggregates (amyloids) can be mentioned. Among different types of probed
materials [35] magnetic nanoparticles of iron oxides show inhibiting and even disaggregating
effect on amyloidal aggregation [36-40].
CONCLUSIONS
In summary, the combined SAXS/SANS analysis complemented by DLS measurements
of magnetoferritin aqueous solutions at loading factors in the interval of 160 – 800 reveals two
competitive effects when increasing the LF. First, a partial disassembling of apoferritin shell in
magnetoferritin, starting from the smallest of the applied LFs is observed. The effect increases
with the LF growth and, in addition to the structure polydispersity (distribution of loading over
the proteins), results in a moderate size polydispersity of magnetoferritin. Second, at LFs above
160 a tendency towards slight aggregation (aggregation number below 10) of magnetoferritin is
observed; it takes place in a wide protein concentration interval of 0.2 – 20 mg ml-1 but is rather
sensitive to the preparation procedure. The aggregation becomes more intensive at LFs above
600.
10
ACKNOWLEDGEMENTS
This work was supported by the Slovak Scientific Grant Agency VEGA (projects No. 0041,
0045), by the European Structural Funds, projects NANOKOP No. 26110230061 and
26220120021, PHYSNET No. 26110230097, PROMATECH No. 26220220186, APVV 0171–
10 (METAMYLC) M-ERA.NET MACOSYS and 226507-NMI3. The kind support from Urs
Gasser at the SANS II instrument (PSI) and Manfred Roessle at the P12 BioSAXS beamline
(EMBL/DESY, Petra III) is acknowledged. This work is based on experiments performed at the
Swiss spallation neutron source SINQ, Paul Scherrer Institut, Villigen, Switzerland. L.A. thanks
the Hungarian Scholarship Board for the support of a short research stay at the IEP SAS, and
acknowledges the financial support from project KMR12-1-2012-0226 granted by the National
Development Agency (NDA) of Hungary.
REFERENCES
[1]
[2]
E.C. Theil, R.K. Behera, T. Tosha, Coord. Chem. Rev. 257 (2013) 579.
N. Galvez, B. Fernandez, E. Valero, P. Sanchez, R. Cuesta, J.M. Dominguez-Vera, C.R.
Chimie 11 (2008) 1207.
[3] M. Ceolin, N. Galvez, P. Sanchez, B. Fernandez, J.M. Dominguez-Vera, Eur. J. Inorg.
Chem. 2008 (2008) 795.
F.C. Meldrum, B.R. Heywood, S. Mann, Science 257 (1992) 522.
[4]
[5] M. Uchida, M.L. Flenniken, M. Allen , D.A. Willits, B.E. Crowley, S. Brumfield, A.F.
Willis, L. Jackiw, M. Jutila, M.J. Young, T. Douglas, J. Am. Chem. Soc. 128 (2006)
16626.
K. Fan, Ch. Cao, Y. Pan, D. Lu, D. Yang, J. Feng, L. Song, M. Liang, X. Yan, Nat.
Nanotechnol. 7 (2012) 459.
K.K.W. Wong, T. Douglas, S. Gider, D.D. Awschalom, S. Mann, Chem. Mater. 10
(1998) 279.
[6]
[7]
[8] M.J. Martinez-Perez, R. de Miguel, C. Carbonera, M. Martinez-Julvez, A. Lostao, C.
[9]
Piquer, C. Gomez-Moreno, J. Bartolome, F. Luis, Nanotechnology 21 (2010) 465707.
Z. Mitroova, L. Melnikova, J. Kovac, M. Timko, P. Kopcansky, Acta Phys. Pol A. 121
(2012) 1318.
[10] O. Kasyutich, A. Sarua, W. Schwarzacher, J. Phys. D: Appl. Phys. 41 (2008) 134022.
[11] D.P.E. Dickson, S.A. Walton, S. Mann, K. Wong, Nanostruct Mater. 9 (1997) 595.
[12] M. Koralewski, J.W. Kłos, M. Baranowski, Z. Mitroova, P. Kopcansky, L. Melnikova,
M. Okuda, W. Schwarzacher, Nanotechnology 23 (2012) 355704.
[13] L.Melnikova, Z.Mitroova, M.Timko,
V.M.
M.V.Avdeev,
Magnetohydrodynamics 48 (2013) 407.
V.I.Petrenko,
J.Kovač, M. Koralewski, M.Pochylski,
P.Kopčanský,
L.Almasy,
Garamus,
[14] L.Melníková, Z.Mitróová, M.Timko, J.Kováč, M.V.Avdeev, V.I.Petrenko, V.M.
Garamus, L.Almásy, P. Kopčanský, Mendeleev Comm. 24 (2014) 80.
[15] M. Koralewski, M. Pochylski, Z. Mitroova, M. Timko, P. Kopcansky, L. Melnikova, J.
Magn. Magn. Mater. 323 (2011) 2413.
[16] M.T. Klem, M. Young, T. Douglas, Mater. Today 8 (2005) 28.
[17] D. Franke, A.G. Kikhney, D.I. Svergun, Nucl. Instrum. Methods A 689 (2012) 52.
[18] P. Strunz, K. Mortensen, S. Janssen, Phys. B: Condens. Matter 350 (2004) E783.
11
[19] U. Keiderling, Appl. Phys. A 74 (2002) S1455.
[20] A.I. Kuklin, A.Kh. Islamov, V.I. Gordeliy, Neutron News 16 (2005) 16.
[21] A.G. Soloviev, T.N. Murugova, A.H. Islamov, A.I. Kuklin, J. Phys. Conf. Ser. 351
(2012) 012027.
[22] W. Häussler, A. Wilk, J. Gapinski, A. Patkowski, J. Chem. Phys., 117 (2002).
[23] A.I. Kuklin, T.N. Murugova, O.I. Ivankov, A.V Rogachev, D.V. Soloviov, Y S Kovalev,
A.V. Ishchenko, A. Zhigunov, T.S. Kurkin, V.I. Gordeliy, J. Physics: Conf. Ser. 351
(2012) 012009.
[24] H.B. Stuhrmann, E.D. Duee, J. Appl. Cryst. 8 (1975) 538.
[25] H.B. Stuhrmann, J. Haas, K. Ibel, M.H.J. Koch, R.R. Crichton, J. Mol. Biol. 100 (1976)
399.
[26] M.V. Avdeev, J. Appl. Cryst. 40 (2007) 56.
[27] M.V. Avdeev, V.L. Aksenov, Phys. Usp. 53 (2010) 971.
[28] O. Glatter, J. Appl. Cryst. 10 (1977) 415.
[29] D.I. Svergun, A.V. Semenyuk, L.A. Feigin, Acta Cryst. A44 (1988) 244.
[30] D. Franke, D.I. Svergun, J. Appl. Cryst. 42 (2009) 342.
[31]
J. Skov Pedersen, Adv. Coll. Inter. Sci. 70 (1997) 171.
[32] C.L.P. Oliveira, B.B. Gerbelli, E.R.T. Silva, F. Nallet, L. Navailles, E.A. Oliveira, J.S.
Pedersen, J. Appl. Cryst. 45 (2012) 1278.
[33] M.V. Avdeev, B. Mucha, K. Lamszus, L. Vekas, V.M. Garamus, A.V. Feoktystov, O.
Marinica, R. Turcu, R. Willumeit, Langmuir 26 (2010) 8503.
[34] M. Kim, Y. Rho, K.S. Jin, B. Ahn, S. Jung, H. Kim, M. Ree, Biomacromolecules 12
(2011) 1629.
[35] W. Wu, X. Sun, Y. Yu, J. Hu, L. Zhao, Q. Liu, Y. Zhao, Y. Li, Biochem. Biophys. Res.
Comm. 373 (2008) 315.
[36] K. Siposova, M. Kubovcikova, Z. Bednarikova, M. Koneracka, V. Zavisova, A.
Antosova, P. Kopcansky, Z. Daxnerova, Z. Gazova, Nanotechnology 23 (2012) 055101.
[37] A. Bellova, E. Bystrenova, M. Koneracka, P. Kopcansky, F. Valle, N. Tomasovicova, M.
Timko, J. Bagelova, F. Biscarini, Z. Gazova, Nanotechnology 21 (2010) 065103.
[38] M. Mahmoudi, F. Quinlan-Pluck, M.P. Monopoli, S. Sheibani, H. Vali, K.A. Dawson, I.
Lynch, ACS Chem. Neurosci. 4 (2013) 475.
[39] L. Xiao, D. Zhao, W.-H. Chan, M.M.F. Choi, H.-W. Li, Biomaterials 31 (2010) 91.
[40] C. Cabaleiro-Lago, F. Quinlan-Pluck, I. Lynch, K.A. Dawson, S. Linse, ACS Chem.
Neurosci. 1 (2010) 279.
12
Fig. 1. Experimental SAXS curves for apoferritin and magnetoferritin with low LF. Smearing
and shift of minima are indicated by arrows against the first minimum. Solid lines correspond to
the model curves obtained by DAMMIF (see text). Relative experimental errors at q < 1.8 nm-1
do not exceed 1%. Inset shows PDD functions of apoferritin and magnetoferritin (results of the
IFT treatment of the experimental curves) and a sphere with radius of 6 nm (model calculations).
Fig. 2. Bead models of apoferritin (a) and magnetoferritin with LF=160 (b) obtained by the
DAMMIF procedure on the data shown in Fig. 1.
13
Fig. 3. SAXS data for magnetoferritin with different LF (intermediate values are covered) and
comparison with apoferritin (LF = 0) and ferritin (LF = 1990): experimental scattering curves (a)
and PDD functions as a result of IFT treatment (b). The relative experimental errors in the
scattering data points at q < 1.8 nm-1 do not exceed 1%. The concentrations of proteins are 2.35
mg∙mL-1 for apoferritin, 44 mg∙mL-1 for ferritin, and 2.81 mg∙mL-1 for magnetoferritin LF160,
7.18 mg∙mL-1 LF260, 6.96 mg∙mL-1 LF350, 7.95 mg∙mL-1 LF410, 8.14 mg∙mL-1 LF430. For
convenient view the curves in (a) are shifted vertically by multiplying by the factor indicated at
the right. In (a) the solid lines show the IFT fits.
14
Fig. 4. LF-dependences of I(0) and Rg found by the IFT treatment of the SAXS experimental
curves in Fig. 3a. The lines are plotted to follow the tendencies. Experimental errors do not
exceed the size of the points.
15
Fig. 5. SANS data for magnetoferritin with different LF (high values are covered) and
comparison with apoferritin (LF = 0) and ferritin (LF = 1990): experimental scattering curves (a)
and PDD functions as a result of the IFT treatment (b). In (a) the relative experimental errors do
not exceed 5 %. The protein concentration in all magnetoferritin solutions based on D2O is 20
mg∙mL-1. The concentrations of proteins are 2.35 mg∙mL-1 in apoferritin and 44 mg∙mL-1 in
ferritin solutions. For convenient view data in (a) are shifted vertically by multiplying by the
indicated factor. In (a) the solid lines show the IFT fits.
16
Fig. 6. LF-dependences of hydrodynamic radius (DLS data) and radius of gyration (SAXS and
SANS data, IFT treatment) of magnetoferritin for different series of samples and comparison
with the corresponding values of apoferritin (Apo) and ferritin. For SAXS and SANS data the
experimental errors do not exceed the size of the points. The lines are plotted to follow the
tendencies.
Fig. 7. Guinier plots for SANS contrast variation data for magnetoferritin solutions with LF 160
(a) and LF 510 (b).
17
Fig. 8. The change in the forward scattered intensity I(0) for two samples of magnetoferritin
(loading factors LF=160 and LF=510) solutions with varying volume fraction of D2O, , in the
solvent. The experimental errors do not exceed the size of the points. Effective match points
corresponding to the intensity minima are indicated by vertical arrows.
18
|
1203.0502 | 2 | 1203 | 2012-10-24T03:33:50 | Identifying influential spreaders and efficiently estimating infection numbers in epidemic models: a walk counting approach | [
"physics.bio-ph",
"cs.SI",
"physics.soc-ph"
] | We introduce a new method to efficiently approximate the number of infections resulting from a given initially-infected node in a network of susceptible individuals. Our approach is based on counting the number of possible infection walks of various lengths to each other node in the network. We analytically study the properties of our method, in particular demonstrating different forms for SIS and SIR disease spreading (e.g. under the SIR model our method counts self-avoiding walks). In comparison to existing methods to infer the spreading efficiency of different nodes in the network (based on degree, k-shell decomposition analysis and different centrality measures), our method directly considers the spreading process and, as such, is unique in providing estimation of actual numbers of infections. Crucially, in simulating infections on various real-world networks with the SIR model, we show that our walks-based method improves the inference of effectiveness of nodes over a wide range of infection rates compared to existing methods. We also analyse the trade-off between estimate accuracy and computational cost, showing that the better accuracy here can still be obtained at a comparable computational cost to other methods. | physics.bio-ph | physics | epl draft
Identifying influential spreaders and efficiently estimating infec-
tion numbers in epidemic models: a walk counting approach
Frank Bauer1 (a) and Joseph T. Lizier1,2
1 Max Planck Institute for Mathematics in the Sciences, Inselstrasse 22, D-04103 Leipzig, Germany
2 CSIRO Information and Communications Technology Centre, PO Box 76, Epping, NSW 1710, Australia
PACS 87.33.Ge -- Dynamics of social networks
PACS 89.75.-k -- Complex networks
PACS 64.60.ah -- Percolation
Abstract -- We introduce a new method to efficiently approximate the number of infections re-
sulting from a given initially-infected node in a network of susceptible individuals. Our approach
is based on counting the number of possible infection walks of various lengths to each other node
in the network. We analytically study the properties of our method, in particular demonstrating
different forms for SIS and SIR disease spreading (e.g. under the SIR model our method counts
self-avoiding walks). In comparison to existing methods to infer the spreading efficiency of differ-
ent nodes in the network (based on degree, k-shell decomposition analysis and different centrality
measures), our method directly considers the spreading process and, as such, is unique in providing
estimation of actual numbers of infections. Crucially, in simulating infections on various real-world
networks with the SIR model, we show that our walks-based method improves the inference of
effectiveness of nodes over a wide range of infection rates compared to existing methods. We also
analyse the trade-off between estimate accuracy and computational cost, showing that the better
accuracy here can still be obtained at a comparable computational cost to other methods.
2
1
0
2
t
c
O
4
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
2
0
5
0
.
3
0
2
1
:
v
i
X
r
a
Introduction. -- Epidemic spreading in biological, so-
cial, and technological networks has recently attracted
much attention (see for instance [1 -- 7]). Most of these
studies focus on the following question: Assume that we
first infect a randomly chosen individual of the network
(patient zero) - how likely is it that a substantial part of
the network will be infected? In these earlier approaches
the network was considered as a whole and the role of pa-
tient zero on the disease spreading process was neglected.
In this letter, we consider the role a single individual
plays in the spreading process rather than the global prop-
erties of the network. It is of particular interest to identify
the most influential spreaders, and to do so without ex-
pensive simulations. This knowledge could, for instance,
be used to prioritise vaccinations for the most influential
spreaders. The number of neighbours of an individual is
a simple but crude approximation for an individual's in-
fluence, and one has to take further topological properties
of the network into account to understand the spreading
process adequately [8, 9]. As such, [5, 8, 9] propose dif-
ferent inference measures for a node's spreading influence
(a)E-mail: [email protected], [email protected]
such as the k-shell decomposition, the local centrality mea-
sure or eigenvector centrality.1 All of these approaches
show strong correlations between their measure of influ-
ence and the (simulated) number of infected nodes. There
is potential for improvement however in:
i. considering
network features encountered by longer infection walks,
and ii. addressing the ultimate goal of estimating the ex-
pected number of infections rather than merely obtaining
correlations. Importantly, one can only predict whether
an infection will be epidemic (i.e. a large portion of the
network will be infected) or harmless from an estimate of
infection numbers, not from correlation scores of an infer-
ence method alone. In this letter, we present an approach
based on counting the number of potential infection walks
from a given initially infected individual. Our method
overcomes the above issues and allows us to consistently
estimate with very good accuracy the expected number of
infections from each patient zero. Moreover, our method
is very efficient and has low computational costs.
1For directed networks other methods exist for ranking the in-
fluence of nodes in different dynamical contexts (e.g.
ranking re-
searchers according to influence on the scientific community [10 -- 12].)
p-1
Frank Bauer et al.
di =(cid:80)
The general model. -- We consider epidemic spread-
ing on network structures. A complex network can be
identified with a graph Γ = (V, E) 2 (here V is the vertex
and E is the edge set) in an obvious way. We say that i
and j are neighbours, in symbols i ∼ j, if they are con-
nected by an edge. In general, we deal with undirected
graphs, though our formulae are trivially extended for the
directed case. Often it is convenient to describe a graph by
its adjacency matrix A = (aij)i,j=1,...,V where the matrix
element aij = 1 if i and j are neighbours and zero oth-
erwise, and V is the number of vertices. Furthermore,
j aij denotes the (out) degree of the vertex i.
a
consider
First we
generalization of
the SIS
(susceptible/infected/susceptible)-model and the SIR
(susceptible/infected/removed)-model.
In our model a
disease is spread in a network through contact between
infected (ill) individuals and susceptible (healthy) indi-
viduals. At a given time step, each infected individual
will infect each of its susceptible neighbours with a given
probability 0 ≤ β ≤ 1 (for simplicity we assume that β is
the same for all pairs of vertices - however generalization
to variable βij is straightforward). An infected individual
will be removed from the network with probability (1− λ)
(modelling either death or full recovery with immunity);
otherwise, an infected individual remains in the network
with probability λ and remains susceptible to (re)infection
at the (very) next time step. For λ = 0 and λ = 1 this
model reduces to the SIR and SIS model, respectively.3
For a given network, we want to find the expected num-
ber of infections given the person that was infected first.
It is natural to think of the spreading process in terms of
infection walks in the corresponding graph. The degree of
a vertex is a first indicator of how many individuals it will
infect, however this neglects all infection walks of length
greater than one - see also [9] where the role of longer infec-
tion walks in epidemic spreading is discussed and numer-
ical simulations were performed.4 Moreover such walks
play a very important role in the dynamics of complex
networks. In the following we will show that it is crucial
to also take longer infection walks into account in order to
get precise results.
The probability p(i, j, k) that vertex j is infected
through a walk of length k given that the infection started
at vertex i can be written as
p(i, j, k) = pinf (i, j, k)psus(i, j, k),
(1)
2For simplicity, we do not allow self-loops or multiple edges.
3 This is slightly different to the usual discrete-time SIS model,
where infected individuals must return to susceptible at the next
time step before reinfection is possible. In our interpretation, all non-
removed nodes are susceptible (i.e. infected and susceptible are not
mutually exclusive). This difference allows us to mathematically
generalise and study smooth transitions from the SIR to the SIS
model, using the walk-counting approach.
4 This study used walk counts from a source node as a predictor
of spreading efficiency. However, unlike the technique we present
it did not convert those counts into a direct estimate of infection
numbers, nor did it consider the appropriate types of walks (i.e. one
must consider self-avoiding walks for the SIR model).
where pinf (i, j, k) is the probability that vertex j is infected
at time k given that vertex j is susceptible at time k, and
psus(i, j, k) is the probability that vertex j is susceptible
(i.e. not removed) at time k, both given that the infec-
tion started at vertex i. (We refer to time here since the
spreading process is updated at discrete time steps; hence
it is equivalent to say that vertex j is infected through a
walk of length k or infected at time k).
For general graph topologies it is difficult and expen-
sive to calculate the p(i, j, k) exactly. In the subsequent
analysis we show how each of pinf (i, j, k) and psus(i, j, k)
in turn can be approximated when we make the following
reasonable simplification assumption: We assume that all
infection walks (of the same as well as different lengths)
are independent of each other, i.e. we treat them as if
they have no edges in common.5 As our simulation re-
sults indicate, this is a reasonable approximation. Using
this independent walk assumption, we approximate:
pinf (i, j, k) ≈ qinf (i, j, k) = 1 − (cid:89)
(1 − p(Pm)).
Pm∈P(i,j,k)
P(i, j, k) is the set of all walks from i to j of length k
and p(Pm) is the probability that the infection takes place
along the walk Pm. This formula is easily obtained by
Pm∈P(i,j,k)(1−p(Pm)) is the probability that
noting that(cid:81)
j is not infected at time k given that it was susceptible
and that the infection started at vertex i. It is insightful
to rewrite the last equation in the following form:
qinf (i, j, k) = 1 − k−1(cid:89)
(1 − λlβk)sk,l
ij
(2)
l=0
where sk,l
is the number of walks from i to j of length k
ij
with l repeated vertices, i.e. the number of walks consist-
ing of k + 1 − l different vertices (including i and j).
Let us have a closer look at the relationship be-
tween pinf (i, j, k) and qinf (i, j, k).
To properly com-
pute pinf (i, j, k), one must compute infection probabili-
ties on each walk Pm in some order, and properly con-
dition these infection probabilities on those of overlap-
ping previously considered walk in {P1, P2, . . . , Pm−1}.
This leads to properly conditioned infection probabilities
p(PmP1, P2, . . . , Pm−1) and the expression:
pinf (i, j, k) = 1 − (cid:89)
Pm∈P(i,j,k)
(1 − p(PmP1, P2, . . . , Pm−1)).
Now,
if infection has not already occurred on one of
these previously considered walks {P1, P2, . . . , Pm−1},
then pinf (i, j, k) only differs from qinf (i, j, k) where Pm
has any overlapping edges with {P1, P2, . . . , Pm−1}. Since
infection did not occur on any of these walks with
shared edges, then some of the shared edges for the
5For walk lengths less than or equal to two this assumption is not
needed since all walks are independent anyway.
p-2
Identifying influential spreaders and estimating infection numbers
walk Pm may in fact already be closed (i.e. dropping
p(PmP1, P2, . . . , Pm−1) below p(Pm)). This yields:
pinf (i, j, k) ≤ qinf (i, j, k) for all k,
(3)
i.e. the independent walk assumption leads to an overes-
timation of pinf (i, j, k).
Let us now study psus(i, j, k) in more detail. We intro-
duce prem(i, j, k) := 1 − psus(i, j, k), (see footnote 3) i.e.
the probability that vertex j is removed at time k given
the infection started at vertex i. For all t ≥ 1, we have:
prem(i, j, t + 1) = prem(i, j, t) + psus(i, j, t)pinf (i, j, t)(1− λ)
=⇒ psus(i, j, t + 1) − psus(i, j, t) = −(1 − λ)p(i, j, t) (4)
Summing (4) over all t from 1 to k − 1 we obtain:
psus(i, j, k) = 1 − (1 − λ)
p(i, j, t),
(5)
t=1
where we used psus(i, j, 1) = 1. As before, we use the in-
dependent walk assumption to approximate psus(i, j, k) by
qsus(i, j, k), and also p(i, j, k) by q(i, j, k) where we define:
q(i, j, k) := qinf (i, j, k)qsus(i, j, k).
(6)
k−1(cid:88)
So by analogy to (5) we define:
qsus(i, j, k) := 1 − (1 − λ)
k−1(cid:88)
t=1
L(cid:88)
(cid:88)
k=1
j
We define the impact of vertex i (the estimated number
of infections given that vertex i was infected first) as
Ii := lim
L→∞ Ii(L) = lim
L→∞
q(i, j, k).
Ii counts the total number of infections, and so is not re-
quired to converge if λ > 0 since then some of the vertices
might be infected several times. Alternatively some other
studies define outbreak size as the number of nodes in-
fected at least once, though this does not inform one as
to whether the infection will die out or not. (Note that
if the nodes cannot be infected several times, i.e. for the
SIR model, both aforementioned quantities coincide.) It
is easy to verify that in considering only walks of length 1,
the impact of vertex i is Ii(L = 1) = βdi. This shows that
the degree di of vertex i is the first order approximation of
the impact Ii. In order to calculate the q(i, j, k) we need
to know all the sk,l
from i to all j
can be completed in O(Dk) steps (average case of counting
walks along homogeneous out-degree D nodes with inde-
pendent edges), and is the computational-time bottleneck
for our method. (We propose later an asymptotically more
efficient calculation for the SIS case). Crucially, while this
asymptotic scaling is the same as for simulating the disease
spreading process, the constant factor of proportionality
for our technique is smaller by several magnitudes.6
ij . The calculation of sk,l
ij
q(i, j, t),
∀k.
(7)
In the following, we restrict ourselves to the special cases
λ = 0, 1 where we obtain the SIR and SIS models.
We observe that qsus satisfies an equation similar to (4):
qsus(i, j, k) = qsus(i, j, k − 1) − (1 − λ)q(i, j, k − 1).
We then consider the connection between psus(i, j, k) and
qsus(i, j, k). In order to investigate this we note that walks
of length one and two always satisfy the independence
assumption. Hence we have pinf (i, j, k) = qinf (i, j, k) and
psus(i, j, k) = qsus(i, j, k) for k = 1, 2.
Now we will prove by induction that psus(i, j, t) ≥
qsus(i, j, t) for all t. First we assume this is true for t = k
(as demonstrated for k = 1, 2 above). Then considering
t = k + 1, combining (1) and (4) we have:
psus(i, j, k + 1) = psus(i, j, k)(1 − (1 − λ)pinf (i, j, k))
≥ qsus(i, j, k)(1 − (1 − λ)qinf (i, j, k))
= qsus(i, j, k + 1),
where we used (3) and our
psus(i, j, k) ≥ qsus(i, j, k). Hence we conclude that:
inductive assumption
psus(i, j, k) ≥ qsus(i, j, k) ∀k.
(8)
That is, we systematically underestimate the probability
psus of being susceptible. Together with the observation
that we systematically overestimate the probability of be-
ing infected in (3), these opposite effects of our indepen-
dent walks assumption may balance each other in (6).
The SIS-model. -- The SIS model corresponds to
the case λ = 1, i.e. where infected nodes always become
susceptible again (i.e. do not die or become immune).
Examples of SIS-type disease spreading include computer
viruses and pests in agriculture where the individuals
(computer/crops) do not develop immunity against the
disease and hence can be re-infected again.
To compute qinf (i, j, k > 1), one has to count the num-
ber sk
ij of different walks of length k between i and j, i.e.
the number of possible infection walks with any number
of repeated vertices l. Crucially, sk
ij is given by the ij-
th entry of the k-th power Ak of the adjacency matrix
A, which is computed in low-order polynomial time, mak-
ing our method asymptotically much more efficient than
simulations for the SIS special case. By equation (2) the
probability p(i, j, k) that j is infected by i through a walk
of length k is then approximated by (with λ = 1):
q(i, j, k) = 1 − (1 − βk)sk
ij ,
(9)
6 The expected number of evaluations e per simulation consists
of D evaluations of disease spread to each neighbour plus the same
expected number of evaluations e per Dβ infected neighbour (on
average, in the sub-critical regime); i.e. e = D + Dβe. One can
solve e = D/(1 − Dβ), but it is more useful to write this to limited
walk length k as O(Dkβk−1). Crucially, we require the number of
repeat simulations to be (cid:29) 1/βk−1 for proper sampling, and new
simulations are required for each β. These two requirements push the
constant factor orders of magnitude beyond that for our technique
since we only need to calculate the SAWs once for all β.
p-3
Frank Bauer et al.
where we used qsus(i, j, k) = 1 since λ = 1 here. We obtain:
I SIS
i
= lim
L→∞
1 − (1 − βk)sk
ij
(10)
L(cid:88)
(cid:88)
k=1
j∈V
(using 00 = 1 by convention if we allow β = 1). Again we
point out that this expression might not converge (partic-
ularly if β is too large) since individuals can be infected
several times. Such divergence has a meaningful interpre-
tation:
i.e. that the infection will remain forever in the
network and will not die out. In (10) we take arbitrarily
long walks into account since the vertices cannot develop
immunity against the disease and so no upper bound for
the maximal length of an infection walk exists.
For other diseases it is more natural to assume that the
vertices can develop immunity after infection.
The SIR-model. -- The SIR model corresponds to
the case λ = 0, i.e. where infected nodes never become sus-
ceptible again after infection as the individuals develop im-
munity or die. As far as infection spreading is concerned,
they are considered removed (i.e. they cannot spread the
virus, nor be reinfected). Examples of SIR-type disease
spreading include most diseases spread among humans.
Since a vertex cannot be infected twice, we have to mod-
ify our previous considerations appropriately. Instead of
general walks we now have to consider self-avoiding walks
(SAWs) or paths. Indeed, it has been previously suggested
that an understanding of SAWs would be useful in epi-
demiology [13], though this was not properly investigated.
It is well known that, compared to counting walks, it is
much more difficult to count SAWs in a graph [14]. How-
ever, instead of explicitly calculating the number of SAWs
one can obtain the number of SAWs recursively [15]. In
particular, the number of SAWs from i to j of length k + 1
is given by:
(cid:88)
sk+1,0
ij
(Γ) =
g:g∼j in Γ
ig (Γ \ j),
sk,0
ig (Γ \ j) is the number of SAWs from i
for k ≥ 1 where sk,0
to a neighbour g of j (in Γ) of length k in the graph that is
obtained from Γ by removing the vertex j. The adjacency
matrix of the graph Γ \ j is obtained from the adjacency
matrix of Γ by deleting the j-th row and column.
Noting that there cannot exist a SAW of length k >
V − 1, we obtain for the overall expected number of in-
fected vertices starting from vertex i (with λ = 0):
V −1(cid:88)
(cid:88)
(cid:16)
k=1
j∈V
(cid:17)(cid:32)
1 − k−1(cid:88)
t=1
(cid:33)
I SIR
i
=
1 − (1 − βk)sk,0
ij
q(i, j, t)
.
(11)
(L) to represent estimates with the sum
We write I SIR
over paths k limited to maximum path length L.
i
Simulation results. -- We provide a brief application
of our technique to simulations of SIR spreading phenom-
ena using: a. the social network structure generated from
email interactions between employees of a university [16]
(giant-component with 1133 nodes and 5451 undirected
edges, diameter 8); b. the structure of the C. elegans neu-
ral network [17, 18] (297 nodes and 2345 directed edges)
to demonstrate a directed network, and c. the collabora-
tion network of the arXiv cond-mat repository [19] (giant-
component with 27519 nodes, 116181 undirected edges,
diameter 16) to demonstrate a larger network.
i
i
For each network, we compute estimates I SIR
(L) from
(11) for maximum (self-avoiding) walk lengths L = 1 to 7
(max. of 5 for the cond-mat network), with variable in-
fection rate β, for each patient zero i. To investigate the
accuracy of these estimates, we also compute numbers of
infections for each patient zero i and β as averages SSIR
over 1000 simulations (10000 for the cond-mat network).
Furthermore, to compare the accuracy of inferences of the
relative effectiveness of each node, we have also measured
the k-shell [8] and eigenvector centrality for each node in
the network (these measures were suggested as useful in-
ferrers of relative spreading efficiency from each node in
[8] and [9]). Using Java code on a 2.0 GHz Intel Xeon
CPU E5-2650, the 10000 simulations SSIR
for all nodes
and β values were completed for the cond-mat network
in around 2000 hours; our estimates I SIR
(L) were com-
pleted for L = 4 and 5 in 30 and 60 minutes respectively;
while with Matlab scripts the degree and eigenvector cen-
trality were completed in less than one second each and
k-shell computed in less than 30 seconds. We note that
computation of the relevant walks sk
ij for SIS models is
significantly more efficient than for SIR, since they can
be directly computed from Ak (as described earlier). We
chose to perform SIR simulations here in order to provide
a greater computational challenge for our technique.
i
i
i
i
i
The extent to which our estimates accurately represent
the relative spreading effectiveness from each patient zero
is examined via the correlation of estimates I SIR
(L) to
simulated results SSIR
for the various networks in Fig. 1,
as well as via their rank order correlations (defined in [9]).
These figures demonstrate that our estimates I SIR
(L)
consistently provide very accurate assessment of relative
spreading effectiveness of the nodes over large ranges of β
for all networks examined, in particular for L > 1 and for
β values in the sub-critical, critical, and the lower-end of
the super-critical spreading regimes. (Critical spreading is
defined as β = βc = 1/αmax [9], where αmax is the largest
eigenvalue of the adjacency matrix A. Fig. 1 indicates βc
and also β−3dB where 30 % of the network is infected on
average (upper super-critical regime) - the number of in-
fections continue to increase very quickly beyond this β).
(L) generally improve as
L increases. Estimates up to only short path lengths L do
not properly capture the effects of spreading on the net-
work structure further away from patient zero when these
nodes become more vulnerable at larger β. In particular,
L = 1 captures only the out-degree of the initially infected
node, and therefore does not represent any network struc-
ture more than one hop away. In general then, one faces
The correlation results for I SIR
i
p-4
Identifying influential spreaders and estimating infection numbers
the network, the structure surrounding each node makes
less of an impact on the spreading efficiency.
(a) Email correlation
(b) Email rank order correlation
(c) C. elegans correlation
(d) C. elegans rank order corr.
(e) Cond-mat correlation
(f) Cond-mat rank order corr.
Fig. 1: Correlations and rank order correlations of various esti-
mates of spreading effectiveness to simulated infection numbers
on the structures of the email interaction, C. elegans neural
and cond-mat networks. Results are plotted for estimates ob-
tained using k-shells (red circles), eigenvector centrality (pur-
ple diamonds), out-degree or I SIR
(L = 1) (blue triangles),
our estimation technique I SIR
(L) using self-avoiding walks of
length L = 2 to 5 for cond-mat and 7 for the other networks
(greyscale ×, darker gray-black to indicate longer walk lengths
L), and for estimates from smaller numbers of simulations (10,
100 and 1000, which increase in accuracy) for the cond-mat net-
work only (green +). Vertical (left) green lines indicate critical
spreading at βc and (right) red lines indicate β−3dB with 30 %
of network infected on average.
i
i
a trade-off between accuracy of inference of effectiveness
against shorter computational time. Importantly though,
very good results can be obtained with short path lengths
L, with the results from L = 4 say being almost indistin-
guishable from longer L for most of the range of β. This is
a crucial point, since the runtime for the computations for
L ≤ 4 is much faster than simulations, and is on the order
of the runtimes for the more simple degree (L = 1) and
k-shell inference methods for the small networks. Finally,
we note that the accuracy of the method drops once β is
well-inside the super-critical regime (even for large L) due
to: i.
insufficient path length at high β, ii. our indepen-
dent walks assumption becomes less valid at high L and
β, and iii. with most nodes infecting a large proportion of
i
i
Further, our estimates I SIR
Crucially, the accuracy achieved by our I SIR
(L) can
only be matched by large numbers of simulations, which
take significantly longer runtime. Fig. 1 shows that, for
the cond-mat network, using only 10 repeat simulations
(with runtime double that of I SIR
(L = 5)) provides much
worse correlations, while comparable correlations up to the
lower super-critical regime can only be achieved with 1000
simulations which cost around 200× more runtime. As
deduced earlier, our technique has asymptotically faster
runtime by a significant constant factor.
(L) are consistently more
accurate with L ≥ 2 than the k-shell inference, for β values
in the sub-critical, critical and early super-critical regimes.
A similar conclusion holds against the eigenvalue central-
ity measure of [9], for all but a couple of β values near
the critical regime in Fig. 1(f); indeed, eigenvector cen-
trality only achieves comparable accuracy near the criti-
cal regime. These are crucial results covering the regimes
of importance (since in the deeper super-critical regime,
a large proportion of the network becomes infected and
understanding the spreading efficiency of various initially
infected nodes becomes less important). Though our esti-
mates are less efficient than k-shell or eigenvalue central-
ity, they are still much faster than simulations, and these
results suggest a strong advantage to using I SIR
(L).
i
i
Additionally, we emphasise that while other tools such
as the k-shell and eigenvector centrality are useful for in-
ferring the relative spreading effectiveness of each node,
they do not actually estimate the number of infections
from each node. This is a distinct feature of our approach
as compared to these other methods. In Fig. 2 we directly
compare our estimates I SIR
(L) to the simulated values
SSIR
for each node i, for several values of β. This clearly
shows that our technique provides reasonable estimates of
the expected number of infections across the sub-critical
spreading regime and up to criticality for L ≥ 4. Indeed,
reasonable accuracy can still be obtained with larger L
into the critical regime, though the time-efficiency bene-
fits of doing so (as compared to simulation) declines.
i
i
i
= I SIR
Fig. 2 also demonstrates quite well the manner in which
estimates improve (in general) with increasing maximum
considered path length L. We see that, while using too
small a maximum path length L appears to be the largest
contributor to the inaccuracy of I SIR
(L) (serving to pull
i
points above the line SSIR
(L)), other errors are in-
troduced by our approximations in (3) and (8) (the former
of which pulls the points below this line by overestimating
the probability of infection). As previously stated though,
for reasonable β and large enough L, these errors seem to
roughly cancel. Importantly also, while the correlation of
estimates to simulated infection numbers may not always
increase with L in the supercritical regime, larger L values
provide consistently more accurate estimates of infection
numbers (e.g. see β = 0.12 in Fig. 1 and Fig. 2).
i
Finally, we consider a simple heuristic to determine an
p-5
0.750.800.850.900.951.00 0.01 0.1Correlationβ0.900.910.920.930.940.950.960.970.980.991.00 0.01 0.1Rank order correlationβ0.700.750.800.850.900.951.00 0.01 0.1Correlationβ0.860.880.900.920.940.960.981.00 0.01 0.1Rank order correlationβ0.600.650.700.750.800.850.900.951.00 0.001 0.01 0.1Correlationβ0.700.750.800.850.900.951.00 0.001 0.01 0.1Rank order correlationβFrank Bauer et al.
(a) β = 0.023
(b) β = 0.040
(c) β = 0.050
(d) β = 0.120
Fig. 2: Simulated SSIR
to 7, using the email interaction network structure. The straight green line represents the ideal plot SSIR
(L) infection numbers, for each patient zero and maximum path lengths L = 1
i
versus estimated I SIR
i
i
= I SIR
i
(L).
appropriate L length to use. For potential infection walks
from patient zero of length L, and for small β in the sub-
critical regime (in particular with doβ < 1), one can make
a naive estimate of the expected numbers of infections at
length L as (doβ)L, where do is the average out-degree of
the network. One can then compute minimum L values
to keep (doβ)L below a given value r. For instance, in the
email network (with do = 9.62), using r = 0.02 suggests
that with L ≥ 4 and β ≤ 0.039 we will only neglect infec-
tions on walk lengths where the expected number of infec-
tions was below 0.02. Of course, this is a simple estimate,
neglecting the effects of dependent walks and making an
implicit assumption that this r = 0.02 is not large enough
to significantly alter the number of infections; however it,
along with observing from the diameters that many walks
can be captured with L = 4, helps to explain why L = 4
provides good results even as β approaches criticality.
Conclusions. -- We have presented a method for effi-
ciently estimating the number of resulting infections from
a given initially infected node in a network model. This
technique focusses on counting the number of functional
walks to each candidate for infection, and in SIR mod-
els the only type of walks of interest are self-avoiding
walks. Our technique is distinct from other recently pro-
posed measures to infer spreading effectiveness of each
node because it focusses specifically on disease spreading
walks rather than general measures of network topology.
We demonstrated our technique to provide consistently
more accurate inference of spreading effectiveness than
other candidate techniques such as k-shells up to the lower
super-critical regime. This accuracy improvement is ob-
tainable in reasonable computational time for SIR models,
while still more accurate assessment can be obtained by
increasing the computational time, and faster assessment
can be made for SIS models. Our technique is also distin-
guished in specifically estimating the number of infections,
and we demonstrated that these estimates are reasonably
accurate over a large range of β for large enough values of
the maximum counted walk lengths L.
∗ ∗ ∗
The research leading to these results has received fund-
ing from the European Research Council under the Euro-
pean Union's Seventh Framework Programme (FP7/2007-
2013) / ERC grant agreement n◦ 267087. We thank the
CSIRO High Performance Computing and Communica-
tions Centre and the Max Planck Institute for Mathemat-
ics in the Sciences for use of their computing clusters.
REFERENCES
[1] Cohen R., Erez K., ben-Avraham D. and Havlin S.,
Phys. Rev. Lett., 86 (2001) 3682.
[2] Albert R., Jeong H. and Barabasi A., Nature, 406
(2000) 378.
[3] Pastor-Satorras R. and Vespignani A., Phys. Rev.
Lett., 86 (2001) 3200.
[4] Dodds P. S. and Watts D. J., Phys. Rev. Lett., 92
(2004) 218701.
[5] Chen D., Lu L., Shang M.-S., Zhang Y.-C. and Zhou
T., Physica A, 391 (2012) 1777.
[6] Newman M. E. J., Phys. Rev. E, 66 (2002) 016128.
[7] Chung F., Horn P. and Lu L., The giant component in
a random subgraph of a given graph Vol. 5427 of Lecture
Notes in Computer Science 2009 pp. 38 -- 49.
[8] Kitsak M., Gallos L., Havlin S., Lijeros F., Much-
nik L., Stanley H. and Makse H., Nature Physics, 6
(2010) 888.
[9] Klemm K., Serrano M. A., Eguiluz V. M. and
San Miguel M., Scientific Reports, 2 (2012) 292.
[10] Lu L., Zhang Y.-C., Yeung C. H. and Zhou T., PLoS
ONE, 6 (2011) e21202.
[11] Zhou Y.-B., Lu L. and Li M., New Journal of Physics,
14 (2012) 033033.
[12] Radicchi F., Fortunato S., Markines B. and Vespig-
nani A., Phys. Rev. E, 80 (2009) 056103.
[13] Eubank S., Jap. J. of Infectious Diseases, 58 (2005) S9.
[14] Madras N. and Slade G., The Self-Avoiding Walk
(Birkhauser) 1993.
[15] Hayes B., American Scientist, 86 (1998) 314+.
[16] Guimer`a R., Danon L., D´ıaz-Guilera A., Giralt F.
and Arenas A., Phys. Rev. E, 68 (2003) 065103+.
[17] White J., Southgate E., Thompson J. and Brenner
S., Phil. Trans. R. Soc. London, 314 (1986) 1.
[18] Watts D. J. and Strogatz S. H., Nature, 393 (1998)
440.
[19] Newman M. E. J., Phys. Rev. E, 69 (2004) 066133.
p-6
0.00.51.01.52.02.53.0 0 0.5 1 1.5 2 2.5 3SiSIR(L)IiSIR(L)L=1L=2L=3L=4L=5L=6L=7 0 2 4 6 81012 0 2 4 6 8 10 12SiSIR(L)IiSIR(L)L=1L=2L=3L=4L=5L=6L=7 0 510152025 0 5 10 15 20 25SiSIR(L)IiSIR(L)L=1L=2L=3L=4L=5L=6L=70100200300400500 0 100 200 300 400 500 600SiSIR(L)IiSIR(L)L=1L=2L=3L=4L=5L=6L=7 |
1201.2607 | 1 | 1201 | 2012-01-12T16:13:16 | Experimental assessment of the contribution of electrodynamic interactions to long-distance recruitment of biomolecular partners: Theoretical basis | [
"physics.bio-ph",
"cond-mat.soft"
] | Highly specific spatiotemporal interactions between cognate molecular partners essentially sustain all biochemical transactions in the living matter. That such an exquisite level of accuracy may result from encountering forces solely driven by thermal diffusive processes is unlikely. Here we propose a yet unexplored strategy to experimentally tackle the long-standing question of a possibly active recruitment at a distance of cognate partners of biomolecular reactions via the action of resonant electrodynamic interactions. We considered two simplified models for a preliminary feasibility investigation of the devised methodology. By taking advantage of advanced experimental techniques nowadays available, we propose to measure the characteristic encounter time scales of dually-interacting biopartners and to compare them with theoretical predictions worked out both in the presence or absence of putative long-range electromagnetic forces. | physics.bio-ph | physics | Experimental assessment of the contribution of electrodynamic
interactions to long-distance recruitment of biomolecular
partners: Theoretical basis
Jordane Preto,1, 2, ∗ Elena Floriani,1, 2, † Ilaria Nardecchia,1, 2, 3, 4, 5, ‡
Pierre Ferrier,1, 3, 4, 5, § and Marco Pettini1, 2, ¶
1Aix-Marseille University, Marseille, France
2CNRS Centre de Physique Th´eorique UMR7332, 13288 Marseille, France
3Centre d'Immunologie de Marseille-Luminy, 13288 Marseille, France
4CNRS, UMR7280, Marseille, France
5INSERM, U1104, Marseille, France
(Dated: May 28, 2018)
Abstract
Highly specific spatiotemporal interactions between cognate molecular partners essentially sus-
tain all biochemical transactions in the living matter. That such an exquisite level of accuracy
may result from encountering forces solely driven by thermal diffusive processes is unlikely. Here
we propose a yet unexplored strategy to experimentally tackle the long-standing question of a pos-
sibly active recruitment at a distance of cognate partners of biomolecular reactions via the action
of resonant electrodynamic interactions. We considered two simplified models for a preliminary
feasibility investigation of the devised methodology. By taking advantage of advanced experimental
techniques nowadays available, we propose to measure the characteristic encounter time scales of
dually-interacting biopartners and to compare them with theoretical predictions worked out both
in the presence or absence of putative long-range electromagnetic forces.
PACS numbers: 87.10.Mn; 87.15.hg; 87.15.R-
2
1
0
2
n
a
J
2
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
0
6
2
.
1
0
2
1
:
v
i
X
r
a
∗Electronic address: [email protected]
†Electronic address: [email protected]
‡Electronic address: [email protected]
§Electronic address: [email protected]
¶Electronic address: [email protected]
1
I.
INTRODUCTION
Living matter hosts a huge number of molecular players (i.e. proteins, nucleic acids) in-
volved in simultaneous yet specific chemical reactions, despite an apparent lack of systematic
spatial order. A phenomenological description of these biomolecular machineries at work
often makes use of the concept of "recruitment", leaving usually unclear how biomolecules
partners encounter or move toward their specific targets and sites of action. On this point
significative progress has been made about DNA-protein interaction at short distance. This
followed the puzzling problem posed by the E.Coli lac repressor-operator protein which was
found to locate its specific DNA-binding site several orders of magnitude faster than the
upper limit estimated for a diffusion-controlled process [1, 2]. A widely accepted approach
to tackle this problem is the so-called facilitated diffusion, on which a vast literature exists
(see for instance Refs.[3–6] and citations therein). To the contrary, for DNA-protein inter-
actions and, more generally, for any dually-interacting biomolecules the mutual approach
from a long distance is addresses to as 3D bulk diffusion and is not further studied (by "long
distance" it is meant: much larger than the Debye screening length). Actually, at first in-
spection, the mutual approach of cognate partners might well be driven by Brownian motion
only, as at living temperature the ubiquitously distributed water molecules move chaotically
in space, colliding with larger/heavier fluid components. On the latter, the total outcome
of many simultaneous hits are forces of both random intensity and direction. Hence, by
displacing themselves in a diffusive way through the inner cellular space, large molecules
sooner or later will encounter their targets.
A complementary proposal, which remains hitherto largely unexplored, is the possibility
for molecules to interact at a distance via the electromagnetic field which is known to have
sizeable magnitudes in living matter [7, 8]. In particular, electromagnetic attractive forces
acting on a long range might, in specific conditions, facilitate the encounters of cognate part-
ners, so that specific biomolecular reactions would occur more effectively than if dependent
on stochastic motion only. Exploring this possibility, it should be stressed that the static
dielectric constant εs of water is particularly high, εs ≃ 80, at physiological temperatures.
In addition to this dielectric screening, freely moving ions in the cellular medium tend to
make the environment electrically neutral; accordingly, the Debye length in a biological
environment is found to be smaller than ≃ 10A, as was estimated on the basis of typical
2
ionic strength of the cytosol [6, 9]. Electrostatic interactions between electrically charged
molecules at a distance larger than the Debye length are very unlikely. Conversely this is
not necessarily the case for electrodynamic interactions [10, 11] since the dielectric constant
depends on the frequency of the electric waves under consideration. Among the latter, the
interactions occurring between oscillating electric dipoles are of a particular interest since in
many cases the long range nature of the interaction potential is essentially "activated" by
the proximity of the dipole frequencies (resonance). In other words, two molecules whose
dipole moments oscillate at the same frequency may undergo a so-called resonant interaction
[12, 13], which is described by the potential U(r) ∝ −1/r3 with r the intermolecular distance
(see Appendix). On the contrary, an off-resonance situation would produce a standard van
der Waals-like potential, i.e. U(r) ∝ −1/r6, typically a short range interaction (see Ap-
pendix). Such a frequency-selective interaction, when applied to a biological context, might
be of utmost relevance during the approach of a molecule toward its specific cognate part-
ner(s). To the best of our knowledge this proposition dates back to Jordan who advanced
the idea that resonant interactions within a quantum framework could play a significant
role in autocatalytic reactions or influence the process of biological synthesis in such a way
that replicas of molecules present in the cell are formed [14]. His theory was questioned
by Pauling [15], who estimated that such forces, supposed to occur only between identical
molecules, could not be large enough to cause a specific attraction between proteins under
the thermal conditions of excitation and perturbation prevailing in living organisms. Other
attempts to explain biological selectivity have been made later on the basis of usual van
der Waals forces [16, 17]. In parallel, in 1968, H. Frohlich proposed a dynamical model [18]
to account for the capacity of biological systems to self-regulate, emphasizing that, under
specific conditions of energy supply to these systems, part of this supply would not be totally
thermalized but would be used to create order in response to environmental perturbations
[19]. In particular, the normal polarization modes of a macromolecule (or of a part of it)
may undergo a condensation phenomenon, characterized by the emerging of the mode of
lowest frequency containing nearly all the energy supply [18]. Then, relying on this model,
Frohlich suggested [20–22] that - when occurring between two biomolecules - such dipole
oscillations could be excited enough to overcome thermal noise leading to the above men-
tioned frequency-dependent forces. Frohlich's seminal work has stimulated many theoretical
investigations until our present days (see for example Refs.
[8, 23–25]). Moreover, a vast
3
literature is available about the experimental observation of low-frequency modes in the
Raman and far infrared (TeraHertz) spectra of proteins [26] and DNA [27]. These spec-
tral features are attributed to collective oscillation modes of the whole molecule (protein or
DNA) or of a substantial fraction of its atoms. A-priori these collective oscillations of the
molecular electric dipole moment could activate the mentioned long-distance attractive and
selective recruitment interactions. However, a clear-cut experimental confirmation of the
existence of the latter ones within a biological context at the molecular level is still lacking.
In the present paper we consider a yet unexplored strategy to experimentally test, at
least in simplified systems, whether these long-range recruitment forces are actually at work
between typical actors of the broad variety of biomolecular reactions in living matter. On
the basis of theoretical computations resorting to elementary and standard methods in the
theory of stochastic processes on the one side, and recent progress on experimental methods
on the other side, we make a first step toward the design of experiments to test whether
such forces are actually at work in living matter.
In Section II we use two dynamical models to highlight qualitative and quantitative
changes between Brownian and non-Brownian encounters of the macromolecular partners of
a generic biochemical reaction. In Section III, we apply our models to the case of attractive
electrodynamic potential U(r) ∝ −1/r3 expected to have effects at long distance, and then
we report the numerical results that have been obtained with realistic parameters. Finally, in
Section IV we discuss how our findings can be used to design an experiment and we conclude
that Fluorescence Cross Correlation Spectroscopy (FCCS) is an appropriate experimental
tool to perform real-time measurements of the association kinetics of dually-interacting
biopartners. It thus seems experimentally feasible to answer the basic questions formulated
above by comparing the outcomes of the prospected experiments versus the theoretically
predicted curves at different concentrations of the reactants.
II. FIRST PASSAGE TIME MODELS
A. Generalities
Our idea is in principle a natural one: different kinds of forces must have different dynam-
ical effects. Thus we attempted to devise an experimental protocol in order to discriminate
4
between the dynamics of purely random encounters between reaction partners versus en-
counters driven by both a stochastic force plus a deterministic long-range force. Then, by
experiments resorting on available techniques, we wondered whether it could be possible to
discriminate between these different dynamical regimes.
A natural way to proceed from the theoretical standpoint, that may be closely related
to experimental as well as physiological conditions, is to consider an aqueous environment,
initially containing NA particles of a species A and NB particles of a species B. Each
molecule A is expected to interact with each molecule B in two ways :
• As soon as the distance between A and B diminishes below a threshold δ, a biochemical
reaction instantaneously takes place, so that the two molecules are not functional
anymore and are considered as out of the system.
• The particle A and the particle B interact at a distance via a two-body potential U(r)
of an electrodynamic type, as long as the two molecules do not get closer than the
distance δ.
From a general point of view, the equations describing the dynamics of the system include
both random and deterministic forces, and therefore may be given in the form:
mA
d2rA,i
dt2 = −γA
mB
d2rB,j
dt2 = −γB
NB
NA
drA,i
dt −
Xj=1
∇AU (rA,i − rB,j) +p2γAkT ξA,i(t)
drB,j
Xi=1
∇BU (rA,i − rB,j) +p2γBkT ξB,j(t),
dt −
(1)
i = 1, . . . , NA, and j = 1, . . . , NB.
Here, mA, mB correspond to the masses, rA,i, rB,j to the positions, and γA, γB to the
friction coefficients of the constituents of each species. T stands for the temperature in
the solution, and k is the Boltzmann constant. ξ(t) is the random process modeling the
fluctuating force due to the collisions with water molecules, usually represented as a Gaussian
white noise process for which hξ α
A,k(t′)i = δαβδikδ(t−t′), where α, β = 1, 2, 3 are related
to each component of the ξA,i's. The same relation is valid also for the ξB,j's.
A,i(t)ξ β
A-priori, equations (1) describe a very complex dynamics, even in the absence of ran-
domness. For example, assuming that the potential is, for each pair of molecules, of the
5
form U(r) = c1/rm − c2/rn, with c1, c2 constants, m, n ∈ N, m > n and n ≤ d (long-range
condition if d is the spatial dimension), the Hamiltonian subset of this system is actually a
nonlinear classical N-body system whose phase space is entirely filled with chaotic trajecto-
ries [28]. At this stage, the addition of random forces may imply that the representative point
of the system nontrivially wanders in phase space, despite the presence of dissipative terms
which, in principle, would generate trivial attractors. Indeed, in the overdamped limit, when
the acceleration terms can be neglected, one is dealing with a randomly perturbed first-order
nonlinear dynamical system which, as integrability is exceptional, is expected to display a
complex (chaotic) dynamics. Nevertheless, instead of undertaking the numerical integration
of Eqs.(1), we decided to look, as a first step, for some analytic result that can be obtained
at the cost of some simplification of the system.
Because the reaction between two particles A and B occurs the first time they come
sufficiently close together, we will have to focus on first passage times of a simplified version
of system (1). Generally, first passage or first return time statistics are difficult to examine
in dynamical systems, then leading one to model the system under study by keeping only
its salient characteristics, either in a deterministic or stochastic manner [29, 30]. Here we
rather choose to still work with equations (1), but to reduce drastically the dimensionality
of the system then leading to keep in the model under study its salient characteristics only.
This was achieved by noting that Eqs. (1) describe the mutual interaction between the two
sets of particles, A and B, but neither the A nor the B particles interact among themselves.
The trajectories of the A particles are indirectly coupled only through the dynamics of the B
particles. Thus, as a first simplifying hypothesis, we assumed that the B particles are fixed,
and as a consequence the dynamical behaviors of the different A particles are independent.
This leads to the decoupling of the individual equations in (1) and hence to the introduc-
tion of a one-dimensional model representative of the generic dynamics of a single A particle.
This model is considered below according to two different versions and solved according to
standard methods [31].
B. Model 1: absorbing plus reflecting boundaries
Let us consider one fixed molecule B located at the position z = 0 and one molecule A,
initially located at z = x (see Figure 1). We first suppose that if A reaches the boundary
6
z = L of the domain, it is reflected back to z < L; whereas when A reaches the position
z = δ for the first time, it is absorbed. The random trajectory z(t) of the molecule A may
be given, as previously, in the form
dz
dt
m
= v ,
dv
dt
= −γv + F (z) +p2γkT ξ(t).
(2)
For times much larger than the characteristic time m/γ, equations (2) will then relax to a
state in which dv/dt → 0. This approximation is justified by the fact that the biomolecules
involved in reactions of interest (protein-protein or DNA-protein) typically weigh thousands
of Daltons, and thus the characteristic relaxation times in aqueous medium are very short.
Therefore, equations (2) for the A particle can be simplified as
dz
dt
=
F (z)
γ
+s 2kT
γ
ξ(t) .
(3)
As it is well known, the one dimensional Langevin initial value problem [31]
dz
dt
= a(z, t) + b(z, t)ξ(t) ,
z(t0) = x,
is equivalent to the Fokker-Planck equation (FPE) for a probability p(z, tx, t0) of finding
the particle at z at time t, given it was at x at t0 ≤ t
∂
∂t
p(z, tx, t0) = −
∂
∂z
[a(z, t)p(z, tx, t0)] +
From (3), one thus obtains
1
2
∂2
∂z2(cid:2)b(z, t)2p(z, tx, t0)(cid:3) .
∂
∂t
p(z, tx, t0) = −
1
γ
∂
∂z
[F (z)p(z, tx, t0)] +
kT
γ
∂2
∂z2 p(z, tx, t0),
which is also known as the Smoluchovski equation.
(4)
(5)
We now look at the time T at which the reaction between A and B occurs. That is,
the first time when particle A reaches z = δ. Since we are considering an absorbing barrier
at z = δ and a reflecting barrier at z = L, the probability P(T ≥ t) and the one that the
particle would still be in the interval [δ, L] at time t are the same
7
FIG. 1: (Color online) A generic initial condition of Model 1 (t = 0). Here x is the initial distance
between A and B; δ is the distance at which A and B react, and L is the position of the reflecting
barrier for A and the position of B is fixed.
P(T ≥ t) =Z L
δ
dz p(z, tx, 0) := G(x, t).
Besides, as F (z) and kT do not explicitly depend on t, p(z, tx, 0), and thus G(x, t), are
homogeneous processes, such that
G(x, t) =Z L
δ
dz p(z, 0x,−t).
(6)
This implies that G(x, t) satisfies the same partial differential equation of p(z, 0x,−t) for
kT
γ
∂2
∂x2 p(z, 0x,−t),
∂x2 G(x, t)(cid:27).
∂2
(7)
z fixed, that is, a backward Fokker-Planck equation
∂
∂t
p(z, 0x,−t) =
1
γ
F (x)
∂
∂x
p(z, 0x,−t) +
leading to
∂
∂t
G(x, t) = −
1
γ(cid:26)F (x)
∂
∂x
G(x, t) − kT
Here, the initial condition p(z, 0x, 0) = δ(x − z) (here δ is the Dirac functional) clearly
gives
G(x, 0) = 1,
if δ ≤ x ≤ L;
and G(x, 0) = 0,
if not,
(8)
whereas the absorbing condition at δ and the reflecting boundary condition at L allow
to write, respectively
G(δ, t) = 0 and
∂
∂x
G(x, t)(cid:12)(cid:12)(cid:12)(cid:12)x=L
8
= 0 ,∀t > 0.
(9)
If one focuses on the mean first passage time τ (x), which represents a characteristic time
scale of the reaction, one has by definition
τ (x) =
∞
Z0
t
∂
∂t
P(T < t)dt = −
∞
Z0
t
∂
∂t
G(x, t) dt =
∞
Z0
G(x, t)dt,
(10)
after integration by parts. Then, by integrating Eq.(7) between t = 0 and t = ∞, and
using the fact that G(x, 0) = 1 and G(x,∞) = 0, we find that τ (x) must satisfy the following
ordinary differential equation
1
γ(cid:26)F (x)
dτ (x)
dx − kT
d2τ (x)
dx2 (cid:27)
−1 =
with boundary conditions τ (δ) = ∂τ (x)/∂xx=L = 0, as it follows from equations (9) and
(10). The solution is found to be [31]
τ (x) = Z x
δ
dy
with
1
ψ(y)Z L
y
dz
γ
kT
ψ(z),
ψ(x) = exp(cid:20)Z x
δ
F (s)
kT
ds(cid:21) = exp(cid:26)−
U(x) − U(δ)
kT
(cid:27) ,
since F (x) = −∂U(x)/∂x. This gives for τ (x):
dy exp(cid:18)U(y)
kT Z x
kT (cid:19)Z L
τ (x) =
γ
y
δ
dz exp(cid:18)−
U(z)
kT (cid:19) .
(11)
(12)
It can easily be checked that the mean first-passage time in presence of an attracting
deterministic potential, generically written as U(x) ∝ −x−n with a given n > 0, is smaller
than the mean first-passage time with Brownian motion only, i.e., when U = 0. Since
exp(cid:0)x−n(cid:1) is a decreasing function of x, we can find an upper limit for the second integral
and thus get
τ (x) <
γ
kT Z x
δ
dy exp(cid:18)−
y
1
kT yn(cid:19)Z L
kT Z x
=
γ
δ
9
kT yn(cid:19)
dz exp(cid:18) 1
dy Z L
y
dz := τ (x)Bwn.
More explicitly
τ (x)Bwn =
γ
2kT (cid:2)(L − δ)2 − (L − x)2(cid:3) =
γ
2kT
(x − δ)(2L − δ − x).
(13)
C. Model 2: two absorbing boundaries
Let us now consider the alternative model where two particles B are fixed at positions
z = 0 and z = l, so that the particle A, initially located at z = x (see Figure 2), is absorbed
as soon as it reaches z = δ or z = l − δ. Such a model is mathematically similar to the
previous one.
FIG. 2: (Color online) A generic initial condition of Model 2 (t = 0). Two molecules B are fixed
at the boundaries x = 0 and x = l; x and l − x are the initial distances between A and the Bs and
δ is the distance at which A and B react.
Simply the deterministic force and the boundary conditions have to be modified. Conse-
quently Eq.((7)) is to be replaced by
∂
∂t
G(x, t) = −
now G(x, t) is defined as G(x, t) =Z l−δ
δ(x − z) gives
δ
1
γ(cid:26)[F (x) − F (l − x)]
∂
∂x
G(x, t) − kT
∂2
∂x2 G(x, t)(cid:27)
(14)
dz p(z, tx, 0). The initial condition p(z, 0x, 0) =
G(x, 0) = 1,
if δ ≤ x ≤ l − δ
and G(x, 0) = 0,
if not,
(15)
and the absorbing boundary conditions give
G(δ, t) = 0 and G(l − δ, t) = 0 ,∀t > 0.
(16)
The mean first-passage time τ (x) defined above, then satisfies the ordinary differential
equation
10
−1 =
1
γ(cid:26)[F (x) − F (l − x)]
dτ (x)
dx − kT
d2τ (x)
dx2 (cid:27) ,
with boundary conditions
The solution is found to be [31]
τ (δ) = τ (l − δ) = 0.
τ (x) = 2(cid:26)Z l−δ
δ
dy
with
δ
1
ψ(y)(cid:27)−1(cid:26)Z x
−Z l−δ
x
dy
dy
x
1
ψ(y)Z l−δ
ψ(y)Z x
1
δ
dw
dw
δ
1
ψ(w)Z w
ψ(w)Z w
1
δ
dz
γ
kT
ψ(z)
dz
γ
kT
ψ(z)(cid:27)
(cid:27)(17)
(18)
dz
φ(y)φ(x)
φ(z)
U(x) − U(l − x) − U(δ) + U(l − δ)
kT
δ
kT
F (s) − F (l − s)
ds(cid:27) = exp(cid:26)−
ψ(x) = exp(cid:26)Z x
since F (x) = −∂U(x)/∂x. After simplification, one has
dyZ l−δ
U(s) − U(l − s)
dy φ(y)(cid:27)−1Z x
γ
kT (cid:26)Z l−δ
τ (x) =
x
δ
δ
φ(s) = exp(cid:26) −
kT
y
dwZ w
(cid:27) .
where
Similarly to the Model 1, the expression for the mean reaction time with Brownian motion
only (U = 0), is particularly simple
τ (x)Bwn =
γ
2kT
(x − δ)(l − δ − x) .
(19)
To summarize, in this Section we have obtained the general form of the mean first-passage
time τ (x), that is the average time needed by molecule A to reach the molecule B (or one
molecule B in the case of Model 2), as a function of the initial intermolecular distance x and
temperature T , for both Model 1 [Eq.(12)] and Model 2 [Eq.(18)], respectively. The same
function is given for randomly driven encounters between the reaction partners in Eq.(13)
for Model 1, and in Eq.(19) for Model 2.
11
III. QUANTITATIVE THEORETICAL PREDICTIONS
In order to answer the question of whether it would be feasible to experimentally detect
the possible existence of a deterministic attractive force through which the cognate partners
of biochemical reactions interact at long distance, we first have to delimit the physical
context, choose the domain of physical parameters and provide the analytic form of the
two-body interaction potential. In what follows, a long range resonant potential potential
U(x) = −C/x3 is considered. As discussed in the Introduction, this kind of interaction
can have sizeable effects at long distances at variance with London - Van der Waals 1/x6
interactions (see also the Appendix). Following Frohlich [22], a lower bound for the coefficient
of this potential is given by C ≃ e2(ZAZB)1/2/2Mω0ε′(ω0), where ZA and ZB denote the
number of charges of averaged mass M and charge e contributing to the dipole moment of the
molecules A and B respectively, whereas ω0 stands for their oscillation frequency; ε′(ω0) is the
real part of the dielectric constant of the interposed medium. In particular, it is interesting
to remark that in the expected range of oscillation frequencies for the setup of collective
dipole oscillations in macromolecules (that Frohlich estimated to be ω0 ≃ 1011− 1012Hz) the
value of ε′(ω0) drops down to a few units [32] thus allowing a much smaller screening of the
interactions with respect to the static case. In this context, we use ZA = ZB = 1000 (see
Ref. [8]) and the proton mass for M. A convenient unit system remains to be chosen. For
the numerical tabulations of τ (x), instead of c.g.s. units we use µm, kDa, and µs, with the
following definitions 1 µm =10−4cm, 1 kDa = 10−21g and 1 µs = 10−6s.
In this system of units we evaluate the lower bound of C which is found to be C ≃
10−30erg.cm3 = 0.1 kDa.µm5.µs−2. Henceforth, we shall consider C varying from 0.1 to
10 kDa.µm5.µs−2. These values are given with a degree of arbitrariness that can be reduced
by considering that C = 10 corresponds to the physical situation where U(x) ≃ kT at
x ∼ 0.1µm. Hence the choice C ∈ [0.1, 10] is a very cautious estimate with respect to those
existing in the literature about a possibly larger range of action (it has been surmised by
Frohlich and others that U(x) might become comparable with kT at x ∼ 1µm or more
[8, 23, 24]). Among other constants appearing in equations (12) and (18), the friction
coefficient γ of the molecule A has been estimated according to Stokes' law γ = 6πη(T )R,
where η(T ) corresponds to the viscosity of water at temperature T and R stands for the
hydrodynamic radius of the molecule. The value of R has been set equal to 5· 10−3µm which
12
is the typical diameter of a biomolecule with a mass in the interval 50 − 100kDa (proteins
and DNA fragments); the same value has been fixed for the reaction radius δ introduced in
both models : R = δ = 5 · 10−3µm.
All the computations of τ (x) have been performed by means of MATLAB programs. Also,
as MATLAB does not allow to perform direct integrations over non-rectangular domains,
integrals with variable limits in equations (12) and (18) have been first "vectorized" [33] for
each x to calculate τ (x) with a recursive adaptive Simpson quadrature (MATLAB quadl
function). Further checks on the reliability of the method have been done through direct
numerical integration of the Langevin equation (3) by means of a standard Euler-Heun algo-
rithm and by averaging over 104 different realizations of the random walk. Minor precision
problems especially when x ∼ δ have been thus detected and corrected in what follows.
A. Model 1
We have computed τ (x) and τ (x)Bwn by means of Eqs.(12) and (13) respectively, where
we have set R = δ = 5· 10−3µm and U(x) = −C/x3, as detailed above. The position L of the
reflecting barrier characteristic of Model 1 has been fixed so as x ≪ L for all x. In particular,
L = 10µm and a maximal value for x equal to 1µm have been used. Figure 3 displays the
numerically found shapes of both functions τ (x) and τ (x)Bwn computed at T = 300K and
for different values of the attractive potential coefficient C. A first check on the reliability
of the plotted results is done by observing that τ (x) < τ (x)Bwn for all the x values while
both curves merge at large x values, as expected when the resonant attraction is wiped out
by thermal noise. In particular, as we always considered x ≪ L, the asymptotic behavior
of τ (x)Bwn is then proportional to x as required by Eq.(13). On the contrary, at smaller
x, τ (x)Bwn bends downwards to slightly smaller values with respect to the extrapolated
linear dependence (this happens when x is no longer much larger than δ). At variance,
the pattern of τ (x) has two asymptotic limiting behaviors: at large x values it joins the
Brownian curve τ (x)Bwn, and at small x values it is τ (x) ∼ x5 (a power-law characteristic
of the 1/x3 form of the potential); the latter might be anticipated on the basis of simple
dimensional arguments since the l.h.s. of Eq.(3) has the dimensions [dz/dt] = lt−1 while the
r.h.s. leads to [C/x4] = [C]l−4 in a purely deterministic regime. By combining the two, for a
generic time scale [τ ] = t associated with a displacement length x we get τ ∼ [x]5 [36]. The
13
two limiting behaviors are bridged by a steep transition pattern which moves rightward or
leftward according to the value of C, as shown on Figure 3; the stronger the potential the
larger the x-values at which τ (x) displays the knee joining the x5 functional dependence.
The transition pattern of τ (x) is steep since the reflecting barrier is located far from the
only one molecule B.
In any case, τ (x)Bwn is found to exceed τ (x) by a factor of, say, 10 at definitely larger
x-values and, what is more relevant, at longer values of first encounter time : with C = 0.1,
such a difference occurs at x ≃ 300A where τ Bwn ≃ 6ms and τ ≃ 600µs, while with C = 1.0,
it occurs at x ≃ 640A where τ Bwn ≃ 10ms and τ ≃ 1ms, and with C = 10 at x ≃ 1400A
where τ Bwn ≃ 20ms and τ ≃ 2ms. As we shall see in the next Section, should we interpret x
as the average distance between any two reacting molecules in three dimensions, this range
of x-values (between a few hundreds Angstroms and 1µm) is easily attained by varying
the concentrations of the reactants between a few micro-Moles down to one nano-Mole.
Notably, the encounter times belong to an interval of values easily accessible by means of
optical detection methods.
A priori, further qualitative indications on the possible presence of attractive deterministic
forces between cognate partners could be observed by modifying the temperature of the
system. As shown in Figure 4, τ (x) and τ (x)Bwn plotted for three different values of T
confirm that the x5 functional dependence is purely deterministic as T has no influence
within this domain. On the contrary, τ (x)Bwn displays the same dependence on T for all
values of x . Surprisingly, the steep transition pattern of τ (x) at intermediate values of x is
characterized by a temperature dependence which is inverted compared to the Brownian case
: in presence of an intermolecular potential, the higher the temperature the larger the first-
passage time of A at x = δ. Finally note that the temperature range considered in Figure
4 is a broad one (T = 200, 300, and 400K). Nevertheless, as the physiological temperature
range corresponds only to a few percent around 300K, it is likely that variations of the first
passage time at different temperatures are too weak to be experimentally detectable within
such an interval. In particular, computations performed for temperature differences of 10K
(typically 290, 300, 310K) with the Model 1 show variations of τ less than five percent of its
value in the Brownian case as well as in the case of Brownian plus deterministic force.
14
FIG. 3: (Color online) Model 1: mean encounter time τ between two molecules A and B, initially
placed at a distance x one from the other. Dotted lines are asymptotic behaviors. Dashed line
refers to purely random encounters. Solid line refers to the combined effect of a random force plus
a deterministic one derived from the potential U (x) = −C/x3. Figure (a) refers to C = 0.1. Figure
(b) refers to C = 10.
15
FIG. 4: (Color) Model 1: temperature dependence of τ (x) for C = 1.0. Red solid and dashed
curves refer to T = 200K, green dashed and solid curves refer to T = 300K, blue dashed and solid
curves refer to T = 400K.
B. Model 2
We have also plotted τ (x) and τ (x)Bwn computed according to Eqs.(18) and (19) respec-
tively, where R = δ = 5 · 10−3µm, l = 2x, and U(x) = −C/x3 as in the case of Model 1.
Figure 5 displays the numerically found shapes of both functions τ (x) and τ (x)Bwn computed
again at T = 300K and for different values of the attractive potential coefficient C.
At first check, the characteristics of τ (x) prevailing for Model 1 (i.e., the x5 functional
dependence at small x, and the tendency of τ (x) to join τ (x)Bwn at large x) also apply
here. Nevertheless, the asymptotic x-dependence of τ (x)Bwn is now proportional to x2 as
obtained from equation Eq.(19) by replacing l by 2x when x ≫ δ. On the other hand, the
absence of a reflecting barrier makes the steep transition feature for τ (x) disappear and be
replaced by a mild crossover at intermediate values of x. This steeper pattern for the first
passage time as a function of x - in the case of purely Brownian diffusion - entails a less
pronounced separation between τ (x) and τ (x)Bwn. In fact, a separation between these two
curves by a factor of 10, (to make the same kind of comparison that we did for Model 1)
16
FIG. 5: (Color online) Model 2: mean encounter time τ between two molecules A and B, initially
placed at a distance x from each other. Dotted lines are asymptotic behaviors. Dashed line refers
to purely random encounters. Solid line refers to the combined effect of a random force plus a
deterministic one derived from the potential U (x) = −C/x3. Figure (a) refers to C = 0.1. Figure
(b) refers to C = 10..
17
FIG. 6: (Color) Model 2: temperature dependence of τ (x) for C = 0.1. Red solid and dashed
curves refer to T = 200K, green dashed and solid curves refer to T = 300K, blue dashed and solid
curves refer to T = 400K.
occurs for C = 0.1 at x ≃ 200A for which τ Bwn ≃ 1.1µs and τ ≃ 0.1µs; while for C = 1.0
this happens at x ≃ 600A for which τ Bwn ≃ 30µs and τ ≃ 3µs; and, finally, for C = 10 at
x ≃ 1170A for which τ Bwn ≃ 120µs and τ ≃ 12µs. Therefore, we can see that Model 2 is
more constraining than Model 1, in the sense that, at equal values of C (that is, at equal
strength of the long range interaction) smaller intermolecular distances and a much faster
tracking of the dynamics of the reactants are needed to discriminate with the same degree
of confidence (arbitrarily set as a factor of 10) between random and non random encounters
of the reactant molecules.
The temperature dependence of both τ (x)Bwn and τ (x) is reported in Figure 6. The
main features of τ (x) in Model 1 are likewise present in Model 2. In addition, we can see
that the inversion in the temperature dependence with respect to the Brownian case (which
was characteristic of the steep transition pattern of Model 1) is no longer there. Thus, the
peculiar temperature dependence of this steep transition pattern could be mostly attributed
to the presence of the reflecting barrier characteristic of Model 1. Likewise in Model 1, we
used again T = 200, 300 and 400K, even though computations carried out at physiological
temperature again yield too weak variations of τ (x) to be experimentally detectable.
18
IV. DISCUSSION AND CONCLUDING REMARKS
The numerical results reported in the preceding Section are in favor of a positive answer
to the main question addressed by the present work. In fact, the numerical study of Models
1 and 2 revealed qualitative differences in the mean first passage time τ between the case
of a pure Brownian diffusion of the molecule A (see Section II) and the case in which an
attractive (resonant) potential U is added to a random force. In particular, in the latter
case, the functional dependence of τ on the initial distance x between the molecule A and
its target (molecule(s) B) demonstrates the existence of different patterns in the two models
depending on the range of the x values considered :
• a deterministic pattern at small x values (small initial separations), characterized by
a power law representative of the potential under consideration (x5 for the resonant
potential used, xp+2 for a general potential of the form U(x) ∝ x−p );
• a Brownian pattern at large initial separations, proportional to x or to x2 depending
upon the symmetry of the system (x for an asymmetric situation as described by
Model 1, and x2 for a symmetric one as described by Model 2);
• a steep transition pattern joining the two asymptotic ones in the case of Model 1, and
a smooth crossover joining the two asymptotic ones in the case of Model 2.
Although complementary computations revealed some interesting features in the tem-
perature dependence of τ , the corresponding degree of variation in a laboratory accessible
interval of temperature is too weak to be experimentally detectable.
In any case, it is obvious that x must constitute an experimentally accessible control
parameter so that the results mentioned above may be used to predict the possible role
of long-range intermolecular forces in biological processes. Notably, such an approach is
not so usual. Indeed, most of the attempts made hitherto in this direction have resulted in
experimental measurements of association constants ka (characteristic of a reaction medium),
which are predictable from the Smoluchowski theory also when intermolecular forces are
considered [34, 35]. The focus of Smoluchovski theory is on the association constant ka
which represents the probability for two molecules to react per time unit, irrespective of their
a = 4πRD ≡ 4πδkT /γ
position. In the case of Brownian encounters, this is given by [4] kB
19
where R is the reaction radius that can be approximated to δ from the current study, and
D is the sum of the diffusion coefficients of the two cognate partners. In the presence of
a = 4πR∗D where R has been replaced by [4, 34, 35]
some interaction potential U, one has k∗
R∗ = R(cid:18) ∞
RR
r−2e−U (r)/kT dr(cid:19)−1
. Now, if ka is experimentally measured for some reaction and
it turns out that kB
a < ka, then this would indicate that some deterministic force is in action
but one can hardly find out the law of the interaction potential because after integration
over r there is no one-to-one correspondence between R∗ (thus k∗
of U(r) [24]. On the other hand, in measuring ka < kB
a , we cannot be sure that the reaction
is simply diffusion-driven because, in this case, chemical times could be long enough to
a) and the functional form
make ka smaller than the corresponding Brownian value. The advantage of our dynamical
approach is that our models still apply by choosing δ as the distance at which A and B get
in contact without reacting, and with the experimental technique discussed below (FCCS)
we can make a distinction between the association time and the chemical times.
In the present situation x values might to some extent be considered in three dimension
as the average distance between two molecular partners A and B, while this quantity can be
easily controlled in laboratory experiments by varying the concentrations of the reactants.
Given the concentrations CA = NA/VA and CB = NB/VB (with VA,B = the initial volumes
and NA,B = the number of molecules of the two species respectively; remark that these num-
bers are controlled through the molarity, i.e. a definite fraction of the Avogadro number),
we get the estimate x = C −1/3
av
for the average intermolecular distance from the average
concentration Cav = (NA + NB)/V , where the reaction volume V = VA + VB. In practice,
as an example, with Cav = 1nM we have x ≃ 1µm as the average distance between any two
molecules, while with Cav = 1µM we have x ≃ 1000A. By working at equimolarity, that is
CA = CB, then C −1/3
is a good estimate of the average distance between one A and one B
av
molecule. Working with nano-Moles of DNA and proteins (enzymes, transcription factors)
is quite standard in molecular biology experiments. With such concentrations of reactants,
both models (1 and 2) predict that the first passage time – that can be interpreted as the
average encounter time between one A and one B molecule initially located at intermolecular
distances of a few thousands of Angstroms – varies in the interval between a few tens of mi-
croseconds to about one millisecond in the presence of an attractive deterministic force that
would sum up to the random force. On the contrary, in the very same conditions, random
20
only driven encounters would exceed the above mentioned encounter times by one or two
orders of magnitude. Again, the distance at which sizeable differences could be observed
may vary significantly depending on the actual value of the resonant potential parameter
C. On the other hand, estimates in literature [8, 23, 24] suggest that these long-range reso-
nance interactions could be effective up to distances in the order of 1µm (the action range
is estimated by computing the distance at which the resonance interaction energy equals
the level of thermal noise kT ). In this respect, in the preceding Section we have limited
ourselves to cautious estimates for the parameter C, focusing on conservative assumptions
for average encounter time varying in the interval 10−5 − 10−3 seconds which can be readily
detected with the aid of Fluorescence Cross-Correlation Spectroscopy (FCCS technology).
This is a powerful technique which is being increasingly applied to the study of diffusion
and chemical reaction rates in complex biological systems using fluorescently labeled macro-
molecules [37–39]. FCCS measures the spontaneous fluctuations of fluorescences δF1(t) and
δF2(t) that arise from the diffusion of fluorescently labeled molecules of type 1 and 2, re-
spectively - illuminated by two laser light beams of different colors - into or out of an open
sampling volume. Even though the size of the detection volume is diffraction limited, the
autocorrelation functions of δF1,2(t) and the cross-correlation function hδF1(t)δF2(t)i can be
altered by processes occurring on smaller spatial scales. These correlation functions provide
information on diffusion properties of fluorescent molecules.
Of course, we are well aware of the fact that the models studied here are simplified de-
scriptions of the reality. Indeed, protein-protein and protein-nucleic acid interactions in vivo
generally take place within complex structural scaffolds such as the membrane cytoskeleton
or the chromatin envelope, which are themselves the subject of highly dynamical regula-
tions (e.g., Refs. [40, 41]); and may also possibly interfere with the spatiotemporal control
of the given reactions (e.g., Refs. [42, 43]). Should resonant electrodynamic interactions be
involved within such an intricate context, it seems illusory at this stage to assess realistic C
values simply based on the proposed experiments. In fact, regarding protein-DNA in vivo
(physiological) interactions for instance, it may well be that the putative values fluctuate de-
pending on a host of variables, possibly including - in a non-mutually exclusive way, charges
on proteins and DNA, the effect of surrounding electrolytes, the nucleic-/amino-acid com-
positions, the length of accessible DNA, etc. However, we stress here that our initial goal, as
described in this article, is to merely probe whether or not biological partners can take ad-
21
vantage, besides thermic diffusion, also of long-distance (0.1-1µm) forces of electrodynamic
origin to eventually interact. If established, this novel concept would then in turn open new
avenues of research to investigate long-standing biological issues, e.g., on the precise defi-
nition of which variables exactly pertain on protein-DNA interactions, and how a diffusing
protein particle may actually recognize the particular cognate DNA site among many other
locations also available. Since we do not expect dramatic qualitative changes out of the
numerical simulations of Eqs.(1) in three dimensions [44], an experimental setup providing
a practical realisation of what has been investigated in the present work could be devised
by resorting - as experimental probes - to three broad classes of interactions: protein-DNA,
protein-RNA, and protein-protein (ligand-receptor). As DNA and RNA molecules have
not a preassigned length, it is implicitly understood that only short fragments are to be
considered (some tens or a few hundreds of base pairs, that is, oligonucleotides or plas-
mides respectively). The proteins interacting with DNA or RNA can be processing enzymes
(helicases, polymerases, recombinases) or transcription factors normally bound at promot-
ers, enhancers, insulators, or silencers. Thus, for example, one could choose two molecular
species consisting, respectively, of a short double stranded DNA molecule (for example a
synthetic oligonucteotide of ∼ 100 base pairs or even less) and a protein with a site specific
affinity for the chosen DNA molecule (i.e., a transcription factor). By labeling the DNA
molecules and proteins with standard fluorophores their dynamical behavior can be followed
by means of FCCS microscopy at different concentrations C = CA = CB of the reactants to
get a characteristic time scale as a function of x = C −1/3. In this way such an experimental
set up should provide - after data fitting - an estimate of the constant C for the resonant
potential considered above. Thus, C = 0 would mean that the reactants meet only under the
action of Brownian diffusion, whereas C 6= 0 would prove the existence at the same time of
the long-range interactions evoked throughout this paper and give quantitative information
about them.
Acknowledgments
We warmly thank V. Calandrini, R. Lima, S. Jaeger and D. Marguet for many fruitful
discussions. Work of the M.P. group has been supported by a BQR grant of the former
University of Aix-Marseille II and by a PEPS grant of the CNRS. Work in the P.F. laboratory
22
is supported by institutional grants from INSERM and CNRS, and by specific grants from
the "Fondation Princesse Grace de Monaco", the "Fondation de France", the "Association
pour la Recherche sur le Cancer" (ARC), the "Fondation pour la Recherche M´edicale"
(FRM), the "Agence Nationale de la Recherche" (ANR), the "Institut National du Cancer"
(INCa), and the Commission of the European Communities.
V. APPENDIX
For the sake of clarity and to help the reader to get a hold of the physical origin of the
U(R) ∝ −1/R3 potential referred to throughout the present work, this Appendix provides
some theoretical elements about the interaction of oscillating electric dipoles.
To begin with, let us recall some basic fact on this subject. Two atoms (or two small
molecules) A and B in their ground states with no net charge excess and vanishing average
dipole moment (i.e. both are nonpolar) interact through the London - Van der Waals
dispersive force. The origin of this interaction is as follows. Though the expectation values
of the dipole operators are zero for nonpolar atoms, quantum fluctuations are responsible
for their instantaneous non vanishing dipolar moments. This entails a non zero dispersion of
dipole moment operator. The energy of the two isolated atoms is corrected at first order by
the dipole-dipole interaction energy which is proportional to the average dipole moments,
thus it vanishes when both atoms are in their ground states.
Instead, the second order
perturbative correction, due to the coupling between instantaneous dipole fluctuations, is
found to be proportional to 1/R6.
(In a QED framework the London - Van der Waals
interaction stems from the exchange of virtual photons between the atoms). This is a short
range potential, so called because the exponent of the power law of R is strictly larger than
3, the dimension of physical space. London - Van der Waals interactions are of generically
weak intensity, whereas they likely become of prime importance in a biological context when
acting at short distances (below the Debye length) together with additional interactions of
chemical type [45].
Remarkably, the first order perturbative correction may be non-vanishing under a degen-
eracy condition. Indeed, if one or both atoms are in an excited state, provided that the
condition for exchange symmetry is fulfilled, that is, they have common eigen-energies in
their spectra, it can be shown [46] that the interaction energy is now proportional to 1/R3,
23
a long range potential.
Interactions of similar kinds to those just mentioned between two atoms (or small
molecules) could exist between macromolecules with an oscillating electric dipole moment.
In this case, the oscillating dipole moment would not be due to the electron motions but,
rather, to conformational vibrations. As already mentioned in the Introduction, this in-
teraction between the oscillating electric dipole moments of reacting macromolecules could
play a relevant role in living matter. In fact, as already quickly recalled in the Introduction,
the high static dielectric constant of water together with the considerable amount of ions
present in living cells tend to screen any electrostatic interaction beyond a distance of a few
Angstroms. However, this electrostatic opaqueness does not hold for an oscillating field: the
higher the frequency of an oscillating field the more transparent an aqueous salted medium.
In fact, the value of the dielectric constant of water at room temperature is a decreasing
function of the frequency [32] and, for example, already at 1THz ε(ω = 1012) ≃ 4; likewise,
the imaginary (dissipative) part of the dielectric constant (which is proportional to the con-
ductivity of the medium due to the presence of free ions) is inversely proportional to the
frequency of the oscillating electric field (according to the Drude equation [48]), so that at
suitably high frequency can be negligible.
Let us a now study the basic mechanism of interaction between two oscillating electric
dipoles before discussing its application to biomolecules [47]. As we show below, these
oscillating dipoles can activate long-range forces that will be shown to be frequency selective.
We consider a one dimensional simplified model in which the dipoles oscillate at frequencies
ωA and ωB respectively. Then a computation of the interaction energy between A and B
can be given which, despite the simplified treatment, allows to grasp some basic physical
facts.
Let µA and µB be the masses of the two oscillators, let their dipole moments be parallel
and given by qZArA and qZBrB, and assume that their mutual separation R is such that
R ≫ rA, rB, then we can write the interaction Hamiltonian as
H =
p 2
A
2µA
+
p 2
B
2µB
+
1
2
µAω2
Ar 2
A +
1
2
µBω2
Br 2
B +
ζq2ZAZB
4πε0R3 rA rB,
(20)
where Zi, i = A, B, stands for an effective number of charges which account for the average
value of the dipole moment of the oscillator i; ζ is a geometrical factor depending on the
orientation of the dipoles with respect to the line joining them (on which the distance R is
24
measured). Then, introducing a mean mass M defined so that µA = MZA and µB = MZB,
the Hamiltonian becomes
H =
1
2M (cid:0)p 2
A + p 2
B(cid:1) +
1
2
Mω2
Ar 2
A +
1
2
Mω2
Br 2
B +
β
R3 rA rB,
(21)
where the transformations (Zi)1/2ri → ri and pi/(Zi)1/2 → pi, i = A, B, have been
introduced (the variables ri and pi are still canonically conjugated) and we put β =
ζq2(ZAZB)1/2/4πε0, where ε0 is the dielectric constant of vacuum - in the absence of a
material medium between the oscillators - to be replaced by ε(ω) when a medium is present.
In matrix form this also reads
H =
1
2M (cid:0)p 2
A + p 2
1
2
B(cid:1) +
M (rA rB)
ω 2
A
β/MR3
β/MR3
ω 2
B
C
{z
rA
rB
.
(22)
}
Matrix C is real and symmetric, thus diagonalizable by means of an orthogonal transfor-
mation. Let ω2
+ and ω2
− the eigenvalues of C (homogeneous to squared frequencies). Under
the action of this transformation the Hamiltonian can be cast in the form of the sum of two
decoupled oscillators, that is,
and it can be easily shown that
+ + p 2
1
H =
2M (cid:0)p 2
√2"(cid:0)ω 2
1
1
2
−(cid:1) +
B(cid:1) ±(cid:26)(ω 2
A + ω 2
ω± =
Mω2
+r 2
+ +
A − ω 2
B )2 +
1
2
Mω2
−r 2
− ,
4β 2
M 2R6(cid:27)1/2#1/2
.
(23)
(24)
By considering rA, rB, pA and pB as observables subject to standard commutation relations,
the energy values of the system are obviously given by
E = ω+(cid:18)n+ +
1
2(cid:19) + ω−(cid:18)n− +
1
2(cid:19)
(25)
where n+, n− ∈ N, that is, are integers. Now, let us consider two opposite physical situations
depending on the relative values of the frequencies ωA and ωB of the oscillators.
1. Consider ωA ≫ ωB (or, equivalently, ωA ≪ ωB), we have
ω± =
A + ω 2
1
√2"(cid:0)ω 2
B(cid:1) ±(cid:0)ω 2
A − ω 2
B(cid:1)(cid:26)1 +
4β 2
A − ω 2
B )2M 2R6(cid:27)1/2#1/2
(ω 2
,
(26)
25
and the denominator of the last term is large enough to give at the lowest order
expansion
B )M 2R6 + ...(cid:21)1/2
ω± =
A + ω 2
1
√2(cid:20)(cid:0)ω 2
√2(cid:20)2ω 2
A,B ±
1
B(cid:1) ± (ω 2
(ω 2
B ) ±
A − ω 2
2β 2
A − ω 2
B )M 2R6 + ...(cid:21)1/2
2β 2
A − ω 2
=
B )M 2R6 + ...
where ωA,B stands for ωA in the computation of ω+ and ωB in the computation of ω−.
= ωA,B ±
2ωA,B(ω 2
(ω 2
β 2
A − ω 2
By substituting this expression in Eq.(25) we get
E = ωA(cid:18)n+ +
1
1
2(cid:19) + ωB(cid:18)n− +
B )M 2R6(cid:26) 1
2(cid:19) +
ωA(cid:18)n+ +
β 2
A − ω 2
2(ω 2
1
2(cid:19) −
1
ωB (cid:18)n− +
1
2(cid:19)(cid:27) + ...
(27)
The first two terms correspond to the unperturbed energies of the oscillators A and
B considered as isolated (R → ∞) while the last term provides the lowest order
correction to the unperturbed energy of the system and due to the interaction, this
interaction potential energy is proportional to R−6. Note that this is functionally the
same as the London - Van der Waals interaction but of a remarkably different physical
origin (real oscillations instead of quantum fluctuations).
2. To the contrary, at resonance, that is, ωA ≃ ωB = ω, the eigen-frequencies (24) are
simply given by
ω± = ωr1 ±
β
Mω 2R3 .
(28)
At long distances (imposed by the reality condition for ω± in this equation) we can
develop ω± near ω and replace such a development into Eq.(25) to obtain
E = ω(n+ + n− + 1) +
β
2MωR3 (n+ − n−) −
β2
8M 2ω3R6 (n+ − n− + 1) + ...
(29)
The first order correction to the energy of the system corresponds to the interaction
energy between the two oscillators at resonance and is proportional to R−3. If both
oscillators are in their ground states, i.e. n+ = n− = 0, the first contribution to
the interaction energy in Eq.(29) vanishes as well as the force given above. The first
non vanishing term is again proportional to R−6. But if the lowest of these modes
26
(ω−) gets more excited than the other (ω+) then the consequence is the activation
of an attractive long-range frequency-selective force. A repulsive force could also be
activated in case n+ > n−.
In the context of Frohlich's theory [8, 18, 49, 50] the above described mechanism of
resonant interaction between oscillating dipoles was surmised to have a great poten-
tial relevance for fundamental biological processes at the molecular level. Frohlich
proposed a model describing the coupling between the elastic vibrations of macro-
molecules and the resulting time variations of their dipole moment; the model predicts
that one or a few Fourier modes of the dipole field oscillation should be strongly (co-
herently) excited provided that the energy supply rate exerted on the macromolecule
by its environment exceeds a threshold value. This energy supply is assumed to de-
pend on the biological activity of the environment (metabolic energy). The strongly
excited mode of oscillation of the molecular dipole moment should be due a collective
oscillation either of the entire molecule or of a subgroup of its atoms. The consequence
of such collective oscillations would be to activate selective long-range recognition and
attraction between cognate macromolecular partners via the above described mech-
anism of resonant interaction. Experimental evidence of the existence of collective
excitations in macromolecules of biological relevance is available for polynucleotides
(DNA and RNA) [27] and for proteins [26] in the Raman and far infrared (TeraHertz)
spectroscopic domains.
[1] A.D. Riggs, S. Bourgeois, and M. Cohn, J. Mol. Biol. 53, 401 (1970).
[2] M.D. Barkley, Biochemistry 20, 3833 (1981).
[3] O.G. Berg, R.B. Winter, and P.H. von Hippel, Biochemistry 20, 6929 (1981).
[4] O.G. Berg, and P.H. von Hippel, Ann. Rev. Biophys. Biophys. Chem 14, 131 (1985).
[5] O.G. Berg, and P.H. von Hippel, J. Biol. Chem. 264, 675 (1989).
[6] A.G. Cherstvy, A.B. Kolomeiski, and A.A. Kornyshev, J. Phys. Chem. B112, 4741 (2008).
[7] See the review paper : M. Cifra, J.Z. Fields and A. Farhadi, Prog. Biophys. Mol. Biol. 105,
223 (2011).
27
[8] J. Pokorn´y and Tsu-Ming Wu, Biophysical Aspects of Coherence and Biological Order,
(Springer, Berlin, 1998).
[9] B. Alberts, D. Bray, J. Lewis, M. Raff, K. Roberts and J. D. Watson, Molecular Biology of
the Cell, (Garland, New York, 1983).
[10] D.P. Craig and T. Thirunamachandran, Molecular Quantum Electrodynamics, (Academic,
London, 1984).
[11] G. Compagno, R. Passante, and F. Persico, Atom-Field Interactions and Dressed Atoms,
(Cambridge University Press, Cambridge, 1995).
[12] M. J. Stephen, J. Chem. Phys. 40, 669 (1964).
[13] A.D. McLachlan, Molecular Phys. 8, 409 (1964).
[14] P. Jordan, Phys. Z. 39, 711 (1938); P. Jordan, Z. Phys. 113, 431 (1939).
[15] L. Pauling, Science 92, 77 (1940).
[16] L. Pauling, Nature 248, 769 (1974).
[17] H. Jehle, Proc. Natl. Acad. Sci. U.S.A. 50, 516 (1963).
[18] H. Frohlich, Int. J. Quantum Chem. 2, 641 (1968).
[19] A typical example of the existence of non thermal behaviors in living matter at the microscopic
level is provided by basic energy conversion mechanisms. According to the estimates provided
by electrochemistry (see, for example, J. Bockris, and S. Khan, Surface Electrochemistry
(Plenum Press, New York, 1993), Chapter 7), the efficiency of energy production in mammals
and humans is very high: about 50%. On the other hand, higher living organisms are at a
temperature T slightly above 300K with an excursion ∆T of a few degrees, whence - according
to the second law of thermodynamics - the thermodynamic (equilibrium) efficiency ∆T /T
should be about 1%, much lower indeed. This is a quantitative example of why fundamental
processes in living matter at the molecular level must stem from a strongly correlated and
coherent dynamics.
[20] H. Frohlich, Phys. Lett. A39, 153 (1972).
[21] H. Frohlich, Proc. Natl. Acad. Sci. U.S.A. 72, 4211 (1975).
[22] H. Frohlich, Advances in Electronics and Electron Physics 53, 85-152 (1980).
[23] S. Rowlands, L.S. Sewchand, R.E. Lovlin, J.S. Beck and E.G. Enns, Phys. Lett. A82, 436
(1981); S. Rowlands, L.S. Sewchand, and E.G. Enns, Phys. Lett. A87, 256 (1982).
[24] R. Paul, R. Chatterjee, J.A. Tuszynski, and O.G. Fritz, J. Theor. Biol. 104, 169 (1983).
28
[25] J. Reimers, L. McKemmish, A. Mark, R. McKenzie, and N. Hush, Proc. Natl. Acad. Sci.
U.S.A. 106, 4219 (2009).
[26] See for instance: P.C. Painter, L.E. Mosher, and C. Rhoads, Biopolymers 21, 1469 (1982);
K.-C. Chou, Biophys. J. 48, 289 (1985); A. Xie, A.F.G. van der Meer, and R.H. Austin, Phys.
Rev. Lett. 88, 018102 (2002), and references quoted in these papers.
[27] P.C. Painter, L.E. Mosher, and C. Rhoads, Biopolymers 20, 243 (1981); H. Urabe, and Y.
Tominaga, Biopolymers 21, 2477 (1982); K.-C. Chou, Biochem. J. 221, 27 (1984); J.W.
Powell, et al., Phys. Rev. A35, 3929 (1987); B.M. Fisher, M. Walther, and P.U. Jepsen, Phys.
Med. Biol. 47, 3807 (2002), and references quoted in these papers.
[28] M. Pettini, Geometry and Topology in Hamiltonian Dynamics and Statistical Mechanics, IAM
Series n. 33, (Springer, New York, 2007).
[29] E. Floriani, R. Lima, Chaos 9, 715 (1999).
[30] E. Floriani, D. Volchenkov, R. Lima, J. Phys. A: Math. Gen. 36, 4771 (2003).
[31] C.W. Gardiner, Handbook of Stochastic Methods, (Springer-Verlag, Berlin, 1985); N.G. Van
Kampen, Stochastic Processes in Physics and Chemistry, (North-Holland, Amsterdam, 1981).
[32] W. J. Ellison, J. Phys. Chem. Ref. Data 36, No. 1, 1 (2007).
[33] L.
Shure,
Two-dimensional
integration
over
a
general
domain
(http://blogs.mathworks.com/loren/2006/04/26/two-dimensional-integration-over-a-general-domain),
MathWorks (2006).
[34] P. Debye, Trans. Electrochem. Soc. 82, 265 (1942).
[35] R.M. Noyes, Prog. React. Kinet. 1, 129 (1961).
[36] Further analysis made in presence of other various potential revealed an asymptotic xn+2
dependence for τ at small x-values when U (x) ∝ −x−n.
[37] S. Maiti, U. Haupts, and W.W. Webb, Proc. Natl. Acad. Sci. U.S.A. 94, 11753 (1997).
[38] E.F. Hom, A.S. Verkman, Biophys. J. 83, 533 (2002).
[39] K. Bacia, P. Schwille, Methods 29, 74 (2003).
[40] B. Mugnier, B. Nal, C. Verthuy, C. Boyer, D. Lam, L. Chasson, V. Nieoullon, G. Chazal, X-J.
Guo, H-T. He, D. Rueff-Juy, A. Alcover, P. Ferrier, PLoS ONE 3, e3467 (2008); A. Pekowska,
T. Benoukraf, P. Ferrier, S. Spicuglia, Genome Res. 20, 1493 (2010).
[41] F. Koch, R. Fenouil, M. Gut, P. Cauchy, T.K. Albert, J. Zacarias-Cabeza, S. Spicuglia, A.L.
de la Chapelle, M. Heidemann, C. Hintermair, D. Eick, I. Gut, P. Ferrier, J.C. Andrau,
29
Nature Struct. Mol. Biol. 18, 956 (2011); A. Pekowska, T. Benoukraf, J. Zacarias-Cabeza, M.
Belhocine, F. Koch, H. Holota, J. Imbert, J.C. Andrau, P. Ferrier, S. Spicuglia, EMBO J.
(2011) Aug 16. doi: 10.1038/emboj.2011.295.
[42] A. Bancaud, et al., EMBO J. 28, 3785 (2009).
[43] A. Chaudhuri, et al., Proc. Natl. Acad. Sci. USA 108, 14825 (2011).
[44] The mean first passage time for Wiener-Einstein processes to attain a given absolute displace-
ment is found to be independent of the dimensionality of the process in: V. Seshadri and K.
Lindenberg, J. Stat. Phys. 22, 69 (1980).
[45] J. N. Israelachvili, Quart. Rev. Biophys. 6, 341 (1974).
[46] H. Margenau, Rev. Mod. Phys. 11, 1 (1939).
[47] A more refined treatment of this problem is given in: J. Preto and M. Pettini, Long range
resonant interactions in biological systems, (2011) preprint.
[48] J.D. Jackson, Classical Electrodynamics,(John Wiley & Sons, New York, 1975).
[49] H. Frohlich, IEEE Trans. Microwave Theor. & Techn. 26, 613 (1978).
[50] H. Frohlich, Rivista Nuovo Cimento 7, 399 (1977).
30
|
1707.05719 | 1 | 1707 | 2017-07-18T16:01:49 | Physiological Aging as an Infinitesimally Ratcheted Random Walk | [
"physics.bio-ph"
] | The distribution of a population throughout the physiological age of the individuals is very relevant information in population studies. It has been modeled by the Langevin and the Fokker- Planck equations. A major problem with these equations is that they allow the physiological age to move back in time. This paper proposes an Infinitesimally ratcheted random walk as a way to solve that problem. Two mathematical representations are proposed. One of them uses a non-local scalar field. The other one is local, but involves a multi-component field of speed states. These two formulations are compared to each other and to the Fokker-Planck equation. The relevant properties are discussed. The dynamics of the mean and variance of the population age resulting from the two proposed formulations are obtained. | physics.bio-ph | physics |
Physiological Aging as an Infinitesimally Ratcheted Random Walk
Institute of Physics, University of Brasilia, 70919-970, Brasilia, DF, Brazil
Bernardo A. Mello∗
The distribution of a population throughout the physiological age of the individuals is very
relevant information in population studies. It has been modeled by the Langevin and the Fokker-
Planck equations. A major problem with these equations is that they allow the physiological age
to move back in time. This paper proposes an Infinitesimally ratcheted random walk as a way to
solve that problem. Two mathematical representations are proposed. One of them uses a non-local
scalar field. The other one is local, but involves a multi-component field of speed states. These
two formulations are compared to each other and to the Fokker-Planck equation. The relevant
properties are discussed. The dynamics of the mean and variance of the population age resulting
from the two proposed formulations are obtained.
Published as: PHYSICAL REVIEW E 82, 021918 (2010)
DOI: 10.1103/PhysRevE.82.021918
PACS numbers: 87.23.Cc, 05.40.Fb, 87.10.Ed, 87.18.Tt
I.
INTRODUCTION
The mathematical study of populations started with
Malthus [1], who, in 1789, declared that the human pop-
ulation grows exponentially. Verhulst realized, in 1838,
that the limited availability of resources prevents that
growth and proposed the now famous logistic equation
[2]. Further progress was made with the predator-prey
equations, independently proposed by Lotka and Volterra
[3, 4] in 1925 and 1926.
In these formulations, each species is described by one
variable that represents the total number of individuals.
It is often necessary to have information about the age
structure of the population. In 1945 Leslie used a matrix
containing the probability of surviving or reproducing
until the next time step [5]. When that matrix is applied
to a vector describing the number of individuals at each
discrete age, the vector with the population at the next
time step is obtained.
McKendrick [6], in 1926, and Von Forest [7], in 1959,
expressed age by using a continuous variable. The age
structure at time t is replaced by the probability density
p(a, t) where a is the chronological age (c-age), i.e., the
time elapsed from the birth of the individuals and the
instant t. The dynamics of a population with a mortality
rate κ(a) is described by the convective equation
changes from one individual to another, Lefkovitch, in
1965, described survival and progression from one class
to the next by a transition matrix, in a Markovian rep-
resentation of aging [8].
The interaction of the individuals with the environ-
ment, among themselves and with other species depends
on the physiological aspects measured by the physiologi-
cal age (p-age, φ), through which they are related to the
c-age. For that reason VanSickle, in 1977, proposed using
the p-age to describe the population [9]. The p-age φ can
be represented, for example, by the individual's weight
or size, whose average increases monotonically with the
c-age a. With the relation between these variables de-
scribed by φ = Φ(a), VanSickle rewrote Eq. (1) as
∂p(φ, t)
∂t
= −
∂v(φ)p(φ, t)
∂φ
− κ(φ)p(φ, t),
(2)
where
v(φ) =
p(φ, t) =
dΦ
da(cid:12)(cid:12)(cid:12)(cid:12)Φ−1(φ)
p(Φ−1(φ), t)
,
, and
(3)
(4)
(5)
v(φ)
κ(φ) = κ(Φ−1(φ)).
∂ p(a, t)
∂t
= −
∂ p(a, t)
∂a
− κ(a)p(a, t).
(1)
tion are the mean value of φ,
Two important measures of the population distribu-
Determining the c-age is not possible in several prac-
tical situations. In these cases, the population is divided
into development classes depending on some character-
istics of the individuals. In spite of the fact that these
classes are related to the c-age, they are not an exact
measurement of it. Since the sojourn time within a class
∗ [email protected]
µ(t) ≡Z φ p(φ, t) dφ,
and the variance of φ
σ2(t) ≡Z (φ − µ)2 p(φ, t) dφ,
(6)
(7)
with p(φ, t) normalized. None of the above approaches
has a parameter that allows to fit the model to the ex-
perimentally measured variance.
Bernardo A. Mello
Preprint -PHYSICAL REVIEW E 82, 021918 (2010)
The variability of the development rate among indi-
viduals was introduced by Lee et al in 1976 [10]. Af-
ter that, much work was done with the individual-based
model of Huston, deAngelis and Post [11]. Kirkpatrick
[12], Clother and Brindley [13] exploited the Ito equa-
tion with a Gaussian noise with mean zero, best known
among physicists as the Langevin equation.
In [14],
Plant and Wilson studied populations with continuous
dynamics within stages and discontinuous stage struc-
tures. A formulation of this model using partial deriva-
tives, the Fokker-Planck equation, was utilized by Buffoni
and Pasquali [15]
∂p(φ, t)
∂t
= −v
∂p(φ, t)
∂φ
+
D
2
∂2p(φ, t)
∂φ2
.
(8)
The time evolution of the p-age depends on several
environment factors. The metabolic rates of insects, for
example, change with the temperature. For that reason
the variables v and D in Eq. (8) could be functions of
the temperature, T . Despite this fact, they will be kept
constant from now on.
A method that has being applied in practical situations
is the calculus of degree-days [16] between the occurrence
of a biofix (a detectable biological event) at time t0 and
the development stage at time t1. Formally it is given by
DD ≡Z t1
t0
(T (t) − T0)H(T (t) − T0) dt,
(9)
where H(x) is the Heaviside step function, T is the envi-
ronment temperature and T0 is the baseline temperature.
Since the development rate is believed to be approxi-
mately proportional to T −T0, the number of degree-days
is proportional to the physiological development. The
peak of an infestation is used as the biofix and the evolu-
tion of the population surge is evaluated by the integral.
It is, notwithstanding, not a population description, but
just an estimate of the development state of a population
sample.
The p-age can also be an abstract indicator of the
individuals' maturity [15]. That description is partic-
ularly suitable for species with well-defined stages [17].
Although that abstract value cannot be determined for
every individual, specific values of the p-age are related
to unmistakable biological events, such as eggs hatching,
emergence from pupae, etc. Since the Fountain of Youth
has not yet been discovered, it makes no sense to allow
p-age to move backward. Even if such backward move-
ments could happen at molecular level, they are forbid-
den at the thermodynamic limit represented by the p-
age of an individual. Therefore, the Langevin and the
Fokker-Planck equations are inappropriate descriptions.
A more suitable mathematical model would be a con-
tinuous variable that moves forward with steps of ran-
dom, positive length. The physical model for such a sys-
tem is an infinitesimally ratcheted random walk (IRRW).
It is not a Brownian ratchet in the sense introduced by
Feynman [18], because, in the present approach: a) back-
ward steps are completely forbidden, not just less prob-
able; b) the ratchet teeth are infinitesimal, meaning that
even the smallest movement advances the ratchet to an-
other tooth; c) no reference is made to mechanics or the
thermodynamics theory.
Two mathematical models for the IRRW are presented
below, disregarding the boundary effects, reproduction,
and death, which results in probability conservation. As
will be shown, in both models the values of µ and σ2
evolve uniformly on time:
dµ(t)
dt
= v
dσ2(t)
dt
= D.
(10)
In each of the models proposed, expressions for v and D
are found.
II. NON-LOCAL FORMULATION
The simplest approach to the IRRW involves the scalar
field p(φ, t). Since we are not concerned with inter-
particle interactions, the equation must include linear
terms only. Furthermore, the dynamics must be trans-
lationally invariant on φ, at least inside a given develop-
ment stage. The most general equation satisfying these
requirements is
∂p(φ, t)
∂t
= −αp(φ, t) + αZ p(φ′, t)f (φ − φ′)dφ′.
(11)
Conservation imposes
Z f (φ)dφ = 1.
(12)
The Fokker-Planck equation corresponds to having
α = 1 and
f (φ) = δ(φ) + vd
δ(φ)
dφ
− D
d2δ(φ)
dφ2
.
(13)
The space-discretized form of Eq. (11) is
dpj(t)
dt
= −αpj + αXk
pk(t)fj−k,
(14)
with fk = f (k∆x)∆x. A form of the Fokker-Planck equa-
tion accurate up to the second order in space can be writ-
ten with fk = 0 for all k /∈ {−1, 0, 1} [19]. On the other
hand, the IRRW implies in fk ≥ 0 for k ≥ 0 and fk = 0
for k < 0. Similarly, it requires
(f (φ) ≥ 0
f (φ) = 0
for φ ≥ 0
for φ < 0
.
(15)
The above conditions exclude the Fokker-Planck equa-
tion or any form of Eq. (8) involving φ derivatives.
2
Bernardo A. Mello
Preprint -PHYSICAL REVIEW E 82, 021918 (2010)
By substituting Eq. (11) in the time derivative of equa-
tions (6) and (7) we can describe the evolution of the
average and variance of the p-age as
dµ(t)
dt
= αhxif ,
dσ2(t)
dt
= αhx2if
(16)
If the above equation are integrated on φ disregarding
the boundary effects, the master equation is found to be
dpi(t)
dt
=Xj
Tijpj(t),
(21)
where
where
hg(x)if =Z g(x)f (x) dx.
(17)
The comparison of these equations with Eq. (10) leads
to the conclusion that α and the function f (x) must be
such that
v = αhxif
D = αhx2if .
(18)
Whatever function f (x) is, if the values of hxif and hx2if
are definite,
it is always possible to rescale f (x) and
choose α so that the above equations are satisfied.
If f (x) doesn't include delta functions centered in 0
then f (x) 6= 0 for values of x 6= 0 and it must, according
to Eq. (18), extends for at least a finite L > 0. Conse-
quently, the IRRW represented by Eq. (11) covers finite
distances in an infinitesimal time interval, meaning that
the particles have infinite speed. Although this fact may
seem strange, it should not be a serious problem, since
even the well-accepted Fokker-Planck equation presents
that property.
Similar use of an integral can be find in [20], where
the usual nonlinear term of the Fisher equation was in-
tegrated over space to express nonlocal competition. In
our model, the integral is over the physiological age, but
it is still referred to as the nonlocal term.
The integral form (11) is not suitable for numerical
calculations, since an integral must be evaluated at each
discrete point φ. Fortunately, there exist functions for
which the integral at φ may be quickly calculated from
the value at φ−∆φ. Two of these functions are the linear
and exponential functions.
III. SPEED STATES FORMULATION
Besides using a non-local formulation and a scalar
field, the IRRW can be described by a multi-component
field p(φ, t) = {p1(φ, t), · · · , pn(φ, t)} with the probabil-
ity density given by
p(φ, t) =Xi
pi(φ, t).
(19)
Each component pi corresponds to a population that
moves without dispersion with speed vi, and switch from
state i to state j with rate Tji. They are subject to a
local dynamic equation
pi(t) ≡Z pi(φ, t) dφ.
(22)
Probability conservation imposes Pi Tij = 0. By sum-
ming Eq. (20) over i we arrive to the dynamics equation
for the population density
∂p(φ, t)
∂t
vi
= −Xi
∂pi(φ, t)
∂φ
.
(23)
Being T a Markov matrix for continuous time,
its eigenvalues are all negative, except by one non-
degenerated eigenvalue λ0 = 0. The components p0
i
of this stationary normalized eigenvector are all greater
than or equal to zero. If λ1 is the second greatest eigen-
value, p(t ≫ 1/λ1) ≈ p0.
Even after the steady state of Eq.
(21) is reached,
the convective term of Eq.
(20) constantly moves the
local population away from the equilibrium. After the
transient is gone, the stationary state p(t) = p0 can be
used to derive the dynamics of
µi(t) ≡
1
p0
i Z φ pi(φ, t) dφ
from Eq. (20), resulting in
dµi(t)
dt
= vi +
= vi +
Tijp0
j µj(t)
T 0
ijµj(t).
1
p0
1
p0
i Xj
i Xj
(24)
(25)
ij ≡ Tijp0
Since T is a continuous time Markov matrix, T 0
j
share the same property. Furthermore, once p0 is an
eigenvector of T with eigenvalue 0, a vector with all el-
ements identical will be an eigenvector of T0 with the
same eigenvalue.
In this stationary regime, the dynamics of
p0
i µi(t)
µ(t) =Xi
(26)
can be derived from Eq. (25) or by integrating Eq. (23):
dpi(φ, t)
dt
= −vi
dpi(φ, t)
dφ
+Xj
Tijpj(φ, t).
(20)
The solution of that equation is
µ(t) = m0 + v0t.
3
dµ(t)
dt
=Xi
vip0
i ≡ v0.
(27)
(28)
Bernardo A. Mello
Preprint -PHYSICAL REVIEW E 82, 021918 (2010)
The unceasing exchange of particles among the speed
states at Eq. (20) connects the population of these states
in bundles moving with speed v0.
The motion of the center of mass of each state depends
not only on its own speed but also on the other states'
centers of mass positions. The motion is governed by the
nonhomogeneous Eq. (25) which possesses the particular
solution
µi(t) = mi + v0t,
with mi satisfying
T 0
ijmj = v0 − vi.
1
p0
i Xj
The homogeneous part of Eq. (25),
(29)
(30)
dµi(t)
dt
=
1
p0
i Xj
T 0
ijµj(t),
(31)
ij/p0
depends on the eigenvalues of T 0
i . The eigenvalues of
T0 are all negative but the zero one. The same property
is shared by T 0
i > 0. The homogenous so-
lution disappears after a while, except for the eigenvector
µi = m0 of the null eigenvalue, which can be included in
mi of eq. (29). Therefore, eq. (29) is the asymptotic
solution of eq. (25).
i because p0
ij/p0
Although the population average position at each
speed state can be ahead or behind the average posi-
tion of the whole group, they all move forward with the
same average speed v0. The null eigenvalue correspond-
ing to the constant eigenvector of T0 means that the
vector mj = const can be added to the solution of (30),
reflecting the translational invariance of the system.
After some algebra, the time evolution of σ2(t) can
be obtained from Eq. (20) and (29) when t ≫ 1/λ1.
Thanks to this result, together with Eq. (27), expressions
for v and D are found
vip0
i ,
v =Xi
D = 2Xi
(mi − m0)vip0
i .
(32)
(33)
The subtraction of m0 from mi removes any dependence
on a uniform translation related to the zero eigenvalue of
T0.
The simplest possible system able to accommodate Eq.
(32) is
k −k(cid:21)
T =(cid:20)−k k
ν(cid:21) .
v =(cid:20)0
The eigenvalues and eigenvectors of T are
λ0 = 0,
p0 =
λ1 = −2k,
p1 =
1
1(cid:21) ,
2(cid:20)1
−1(cid:21) .
2(cid:20) 1
1
(34)
(35)
(36)
4
When these values are substituted in Eq. (30) the values
of mi are obtained (the two equalities of the linear system
are not linearly independent)
m =(cid:20)m0 − ν/4k
m0 + ν/4k(cid:21) .
(37)
The average value of φ of the whole population in
t = 0 would be m0, provided that equilibrium was al-
ready reached at that time. The average position of the
zero speed individuals is behind the average population's
position, while the ν speed individuals move ahead of the
group.
Concluding, the constants for the two speed states sys-
tem are
v =
ν
2
D =
ν2
4k
.
(38)
IV. COMPARING THE MODELS
The time evolution of µ(t) and σ2(t) obtained by nu-
meric integration of the three models can be seen in fig-
ures 1a and 1b. When t <
∼ 1, σ2(t) of the two speeds
model slightly deviates from the linear behavior, which
may be due to the transients of the homogeneous solution
of Eq. (25). Except for that, these two curves perfectly
agree with the analytic results.
Figure 2a shows that the three approaches result
in very different population distribution for short time
scales.
In the Fokker-Plank dynamics the initial delta
distribution becomes immediately Gaussian, while the
disappearing of the delta takes some time in the other
two models. Aiming to a quantitative evaluation of this
phenomenon, we define the quantity
q(t) =Z δ(φ) p(φ, t) dφ.
(39)
The evolution of q(t) can be seen in figure 1c.
The instantaneous disappearing of the delta function in
the Fokker-Plank equation, resulting in q(0) ≈ 0, is only
possible due to the singularity of the laplacian of the delta
function. The nonexistence of a similar singularity in the
non-local model prevent the same effect in this model. In
the two speeds model, the singular derivative of dp/dφ in
the v > 0 state instantly moves the population at this
state away from φ = 0, resulting in q(0) = 1/2.
If we apply the frame of reference change x → x− νt in
the two states model, the only effect would be replacing
v in eq. (34) by
0 (cid:21) .
v =(cid:20)−ν
(40)
The resulting distribution with such v would be trans-
lations of the figure 2 distributions. On the other hand,
symmetry analysis implies that the distributions result-
ing from eq. (34) or from eq. (40) should be symmetric
Bernardo A. Mello
Preprint -PHYSICAL REVIEW E 82, 021918 (2010)
(a)
(b)
Gaussian
Exponential
Two states
0
0.5
1
1.5
t
2
2.5
3
(c)
µ
3
2
1
0
2
σ
q
0.2
0.1
0
1
0.8
0.6
0.4
0.2
0
0
0.1
0.2
t
0.3
0.4
0.5
FIG. 1. Results of the numerical integration according to
the Fokker-Plank equation (Gaussian), non-local IRRW with
exponential function and two speed states model, with v = 1
and D = 0.1. The initial distribution was p(φ, 0) = δ(φ), with
the population of the two speed model already in stead state:
p0(t = 0) = p1(t = 0). a) Population average position. b)
Population position variance. c) Value of q(t) as defined in
Eq. (39).
by reflection, and a delta should be present at the right of
the bells seen in figures 2a and 2b. A trace of such delta
function is present on figure 2a, but complectly disap-
peared on figure 2b, due to the numerical discretization.
The asymmetry of the distribution resulting from the
non-local model is visible when we compare that distribu-
tion, in figure 2, with the perfectly symmetric the Fokker-
Planck distribution. That asymmetry is not a surprise,
since the non-local Eq. (11) is not symmetric.
After some time, the distribution at φ > 0 of the IRRW
models look like a truncated bell, but the curves become
bell shaped only after the distribution moves far enough
from the initial position. The two states model becomes
Gaussian faster than the non-local model. This happens
despite the initial asymmetry generated in the two states
model by the numerical error.
Regarding numerical
the Fokker-
Planck equation usually involves inverting a tri-diagonal
implementation,
Gaussian
Exponential
Two states
0
0.5
φ
1
1.5
2
10
)
φ
(
p
)
φ
(
p
8
6
4
2
0
2
1
0
1
(a)
t = 0.1
(b)
t = 0.3
(c)
t = 1
)
φ
(
p
0.5
0
-0.5
0.8
0.6
(d)
t = 3
)
φ
(
p
0.4
1
1.5
2
2.5
3.5
4
4.5
5
3
φ
(e)
t = 10
0.2
0
0.4
0.3
0.2
0.1
0
)
φ
(
p
6
7
8
9
10
φ
11
12
13
14
FIG. 2. Population distribution resulting from the same nu-
merical integration of figure 1 at different instants for the
three models.
5
Bernardo A. Mello
Preprint -PHYSICAL REVIEW E 82, 021918 (2010)
matrix when using second-order accuracy on time. The
two speeds model methods doesn't, since it can be im-
plemented using upwind differencing.
When applying the boundary conditions to the Fokker-
Planck equation, it is necessary to prevent the movement
in the wrong direction. One doesn't need to worry about
such matters when using the IRRW models. On the other
hand, it is awkward to include far-reaching effects of the
non-local formulation in the boundary conditions, mainly
between development stages.
Except for involving more than one field, the numerical
implementation of the two speed states model is overall
more convenient than the Fokker-Planck and the non-
local models.
Only specially planned experiments, or a deeper un-
derstanding of the aging process, will decide which, if
any, of the three models discussed in this work is correct.
Whatever the answer is, it won't invalidate the use of
the other methods as a convenient approximation, not to
mention the fact that many different behaviors can be
accommodated in the freedom of choosing f (x) in Eq.
(11) or vi and Tij in Eq. (20).
V. CONCLUSION
Two basic equations describing the evolution of the
population distribution were proposed. As already men-
tioned, the main problem with the Fokker-Planck equa-
tion is its non-physical properties, namely, negative vari-
ation of φ and the infinite speed. The non-local formu-
lation solves the first problem but not the second, while
the speed states formulation solves both of them.
An advantage of the non-local and the two speed mod-
els over the Fokker-Plank equation is their convenience
for certain numerical implementations. The analytical
expressions for the mean and the variance of the p-age
help to understand the role played by the parameters of
the dynamic equations.
Several other biological phenomena could be included
in a more detailed formulation. Some of them, such as
the age structure, the quiescence, and the dependence
of the biological development on the temperature, could
be included by making the equation parameters depen-
dent on temperature and age. Other phenomena such as
death, reproduction, and diapause, require the introduc-
tion of new terms in the dynamic equation. Introducing
more complex effects such as spatiality and inter species
interaction can only be done by defining new independent
variables and fields.
Real situations demand taking into account some of
these phenomena, requiring several additional assump-
tions and the determination of the corresponding param-
eters. Possible usages include plague control, epidemics,
ecological management, demography, etc. They go well
beyond the scope of this paper, which presents the equa-
tions ruling the intra-stage development.
Most models of genetic agents explored by physicists
use c-age, either discrete [21] or continuous [22]. These
models could be extended by incorporating the p-age,
particularly, the IRRW presented here.
ACKNOWLEDGMENTS
I am grateful to Fernando A. Oliveira for the thought-
ful discussions and to the National Council for Scientific
and Technologic Development - CNPq, for the financial
support.
[1] T. R. Malthus, An Essay on the Principle of Population
(2000)
(J Johnson, London, 1798)
[14] R. E. Plant and L. T. Wilson, J. Math. Biol 23, 247
[2] P. F. Verhulst, Correspondance Math´ematique
et
(1986)
Physique 10, 113 (1838)
[3] A. J. Lotka, Elements of physical biology (Williams and
Wilkins, Baltimore, 1926)
[4] V. Voltera, Nature 118, 558 (1926)
[5] P. H. Leslie, Biometrika 33, 183 (1945)
[6] A. G. McKendrick, Proc. Edinburgh Mat. Soc. 44, 1
(1925-6)
[7] H. von Foerster, "The kinetics of cellular proliferation,"
(Grune and Stratten, New York, 1959) Chap. Some re-
marks on changing populations, pp. 382 -- 407
[8] L. P. Lefkovitch, Biometrics 21, 1 (1965)
[9] J. VanSickle, J. Theor. Biol. 64, 571 (1977)
[10] K. Y. Lee, R. O. Barr, S. H. Gage, and A. N. Kharkar,
J. Theor. Biol. 59, 33 (1976)
[11] M. Huston, D. DeAngelis, and W. Post, BioScience 38,
682 (1988)
[12] M. Kirkpatrick, Ecology 65, 1874 (1984)
[13] D. R. Clother and J. Brindley, Bull. Math. Biol. 62, 1003
[15] G. Buffoni and S. Pasquali, J. Math. Biol. 54, 555 (2007)
[16] W. J. Roltsch, F. G. Zalom, A. J. Strawn, J. F. Strand,
and M. J. Pitcairn, Int J. Biometeorol 42, 169 (March
1999)
[17] R. M. Nisbet and W. S. C. Gurney, Theor. Popul. Biol.
23, 114 (1983)
[18] R. P. Feyman, R. B. Leighton, and M. Sands, "The Fey-
(Addison-Wesley, Reading,
man lectures on physics,"
1963) p. 46.1
[19] It is also possible to write an equation where fk is dif-
ferent of zero only when k ∈ 0, 1, 2. However, this for-
mulation involve a negative f1 which, besides having no
probabilistic interpretation leads to numerical instabili-
ties.
[20] J. A. R. da Cunha, A. L. A. Penna, M. H. Vainstein,
R. Morgado, and F. A. Oliveira, Physics Letters A 373,
661 (2009)
[21] T. J. P. Penna, J. Stat. Phys. 78, 1629 (1995)
6
Bernardo A. Mello
Preprint -PHYSICAL REVIEW E 82, 021918 (2010)
[22] W. Hwang, P. L. Krapivsky, and S. Redner, Phys. Rev.
Lett. 83, 1251 (1999)
7
|
1508.01594 | 1 | 1508 | 2015-08-07T03:20:51 | Hydrodynamics of stratified epithelium: steady state and linearized dynamics | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.TO"
] | A theoretical model for stratified epithelium is presented. The viscoelastic properties of the tissue is assumed to be dependent on the spatial distribution of proliferative and differentiated cells. Based on this assumption, a hydrodynamic description for tissue dynamics at long-wavelength, long-time limit is developed, and the analysis reveals important insight for the dynamics of an epithelium close to its steady state. When the proliferative cells occupy a thin region close to the basal membrane, the relaxation rate towards the steady state is enhanced by cell division and cell apoptosis. On the other hand, when the region where proliferative cells reside becomes sufficiently thick, a flow induced by cell apoptosis close to the apical surface could enhance small perturbations. This destabilizing mechanism is general for continuous self-renewal multi-layered tissues, it could be related to the origin of certain tissue morphology and developing pattern. | physics.bio-ph | physics | Hydrodynamics of stratified epithelium: steady state and
linearized dynamics
Wei-Ting Yeh1,2 a Hsuan-Yi Chen1,2,3 b
1Department of Physics, National Central University, Jhongli, 32001, Taiwan
2Institute of Physics, Academia Sinica, Taipei, 11529, Taiwan
3Physics Division, National Center for Theoretical Sciences, Hsinchu, 30013, Taiwan
(Dated: August 13, 2018)
Abstract
A theoretical model for stratified epithelium is presented. The viscoelastic properties of the tissue
is assumed to be dependent on the spatial distribution of proliferative and differentiated cells.
Based on this assumption, a hydrodynamic description for tissue dynamics at long-wavelength,
long-time limit is developed, and the analysis reveals important insight for the dynamics of an
epithelium close to its steady state. When the proliferative cells occupy a thin region close to the
basal membrane, the relaxation rate towards the steady state is enhanced by cell division and cell
apoptosis. On the other hand, when the region where proliferative cells reside becomes sufficiently
thick, a flow induced by cell apoptosis close to the apical surface could enhance small perturbations.
This destabilizing mechanism is general for continuous self-renewal multi-layered tissues, it could
be related to the origin of certain tissue morphology and developing pattern.
PACS numbers: 87.17.Ee, 87.15.La, 87.19.R
5
1
0
2
g
u
A
7
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
4
9
5
1
0
.
8
0
5
1
:
v
i
X
r
a
a [email protected]
b [email protected]
1
I.
INTRODUCTION
Biological tissues are viscoelastic. Concepts developed for describing passive viscoelastic
materials have been applied to biological tissues [1, 2]. However, active processes such as
forces generated by the cells, cell movement, and cell proliferation make biological tissues
different from passive viscoelastic materials. Thus biological tissues fall into the category of
active soft matters [3]. On the other hand, junctions between neighboring cells prevent free
diffusion of cells in a tissue. It makes biological tissues different from other active biological
systems such as bacterial colonies and cell cytoplasm. As a result, biological tissues become
interesting systems to study complex collective motion and morphogenesis. For example,
it is known that the homeostasis state of a tissue often has undulating surface to increase
the contact area with its environment. Understanding the formation and maintenance of
such spatial structure is thus important for fundamental research of active matters and for
applied research such as fabricating functional surfaces and flexible electronics [4].
Experiments have demonstrated that at short time scales a tissue is solid-like due to the
adhesion junctions between cells [5]. Since the residual stresses can relax due to turnover of
junction proteins [2, 6] and cell rearrangement induced by cell division and cell apoptosis [7],
a tissue is liquid-like at long time scales [1, 2, 8 -- 11]. This suggests a characteristic time τ
such that at time scales smaller than τ a tissue behaves like a solid with shear modulus E.
On the other hand, at time scales larger than τ , a tissue behaves like a fluid with viscosity
η. Typically E ∼ 102 − 104 Pa [10], and τ ∼ 10 − 102 s [8, 10], therefore η = Eτ is on the
order of 103 − 106 Pa s [12].
In this study we focus on stratified epithelium, a multi-layered continuous self-renewal
tissue [13]. Epithelium often forms the outermost layer of skin and mucous membrane, and
acts as a barrier separating the outside and inside of a multicellular organism [14]. On the
top of a stratified epithelium is a free apical surface, on the bottom is a basement membrane
attached to the underlying connective tissue composed primarily of collagen and elastin
fibers. Typically the thickness of stratified epithelium is ∼ 100 µm and the thickness of
the underlying connective tissue is ∼ 1 mm [15]. The proliferative cells are often localized
near the basal lamina, suggesting the existence of a special microenvironment called stem
cell niche [16, 17]. Above the proliferative cells are terminally differentiated cells. They are
functional cells that do not divide, instead they undergo programed cell death (apoptosis).
2
Intuitively, different cells should have different mechanical properties. As supported by
recent experiments, stem cells and tumor cells often have smaller stiffness than normal
differentiated cells [18, 19]. This suggests that the mechanical properties of a tissue should
also depend on local tissue composition [7]. The effects of such inhomogeneity on tissue
dynamics is the focus of this study.
Undulating/fingering structures that are important for biological functions are often seen
in the apical surface or basal membrane of a stratified epithelium. For example, skin wrinkles
or folds are formed when a skin is deformed by external/residual stresses with wavelength
∼ 1 mm [15, 20 -- 22]. At length scale ∼ 100 µm, rete pegs interdigitate with dermal papillae
are observed in the interface between skin epithelium and stroma [23 -- 25]. Similar structures
have also been observed in the epithelium of mucous membrane [20, 26 -- 28]. Studies have
shown that elastic forces could be important during the formation of the aformentioned
structures. By modeling a stratified epithelium as an elastic solid in contact with the un-
derlying elastic connective tissue, the formation of skin wrinkles was associated with the
buckling instability originated from the competition between bending energy of epithelium
and stretching energy of connective tissue [15, 21]. The rete peg structure was suggested to
arise from the incompatible growth of epithelium and connective tissue [29]. Similar mech-
anism can also explain the formation of fingerprint [30], crypt [31], and surface wrinkling
of tumor spheroids [32]. On the other hand, hydrodynamic instabilities are also likely to
produce undulating morphologies in stratified epithelium. For example the formation of rete
peg structure could arise from proliferation-induced stresses [33], and an essential condition
of this result is that the net cell proliferative rate deceases with the distance from basal
lamina. Some models considered more details such as cell lineage dynamics [34], viscoelas-
ticity of the underlying stroma [35], and the effect of growth factors inside a tissue [34, 35].
Overall, several interesting structures and patterns can be understood by a continuum de-
scription of tissues without describing the movement of individual cells and biomolecular
events in details. The mechanical stresses induced by cell turnover inside a tissue [7, 12] is
sufficient to do this job.
The aim of this work is to study theoretically how inhomogeneity of mechanical properties
in a tissue affects the tissue dynamics. For illustrative purpose we consider a stratified
epithelium on a rigid stroma. We begin with a simple model in which the mechanical
properties of the tissue is assumed to be homogeneous. With proliferative cells occupying
3
the lower part, and terminal differentiated cells occupying the upper part of the tissue, this
model has a homeostasis state with a flat apical surface. The linearized dynamics close to
the homeostasis state in this model already reveals interesting effects from flows and cell
proliferation. For example, cell division/apoptosis speed up tissue relaxation towards the
homeostasis state; but this effect is relatively weak when the wavelength of the perturbation
is of the order of the thickness of the tissue.
Due to frequent cell division events, tissue viscosity in the region rich in proliferative
cells should be different from the rest of the tissue. This feature is included in our second
model. Such a modification indeed leads to a different dispersion relation for tissue dynamics
close to the homeostasis state. Especially, when the proliferative region becomes sufficiently
thick, cell apoptosis close to the apical region could induce a flow that strongly hinders the
relaxation towards the steady state. This effect is especially significant for perturbations with
wavelength comparable to tissue thickness (∼ 100 µm for typical stratified epithlium). Since
our model is quite general, we expect this destabilizing mechanism to exist in continuous
self-renewal tissue, but it is not easy to be observed in normal healthy tissue because the
proliferative cells only occupy a thin region. However, during development or regeneration
there is often a transient increase of proliferative cell inside tissues [36]. As a result, the new
effect described by our second model could happen, and it is even possible that undulating
apical surface could be generated by this mechanism.
This article is organized as follows. Both models are presented in Sec. II, where we
also discuss the homeostasis states and linearized dynamics close to the steady states. In
Sec. III we discuss the flow in an epithelium that is induced by cell division/apoptosis. Using
the result derived in the Appendix, we also discuss the relation between the dimensionless
parameter r = rD/rS and ηrel = ηD/ηS, where rD and rS are cell apoptosis rate and cell
division rate, and ηD, ηS are viscosity in region rich in differentiated cells and proliferative
cells, respectively. In Sec. IV, based on these results, we make a conclusion and propose
possible experimental tests to validate our theory. The Appendix presents the relaxation
rate in our second model, then a derivation of the tissue viscosity from a hydrodynamic
model is presented.
4
II. MODELS
A. First model: stratified epithelium with constant viscosity
Consider a basel lamina sitting on the xy-plane. An epithelium grows from this surface
into z > 0 region. On top of the epithelium is a free apical surface. We are interested
in the steady state of this system and the linearized dynamics close to this steady state.
Since epithelial growth typically takes several days to complete [23, 24], in this article we
focus on the long time behavior of our model. Therefore processes with relaxation time not
longer than the time scale τ introduced in Sec. I have all decayed away. The elasticity of
the tissue can be neglected, and the tissue behaves as a viscous fluid [8 -- 11]. The viscosity
of the tissue depends on cell-turnover events [7] and the rearrangement of junction proteins
[2, 6]. Experimentally, the viscosity of a tissue can be estimated from the elastic constant
E and relaxation time τ through the relation η = Eτ .
Since the dynamics is slow, the inertia of the tissue is also negligible. Imposing incompre-
sibility for simplicity, the equation of continuity, force balance, and consitutitive equation
are
∂lvl = kp,
∂iσik = 0,
σik = −pδik + 2ηvik.
(1)
(2)
(3)
Here vi is the i-th component of the flow velocity field, vik ≡ (∂ivk + ∂kvi)/2 is the strain
rate tensor in the tissue, p is the tissue pressure, η is the viscosity of the tissue, and kp is the
(net) cell proliferation rate in the tissue. In general kp is a decreasing function of z because
both nutrient and proliferative cells are relatively abundant near the basal lamina. In this
model η is chosen to be a constant for simplicity. However, in general η should depend on
local tissue composition through forces generated during cell division and cell apoptosis [7],
this will be considered in Sec. II B when we introduce our second model.
We focus on the situation when the tissue is soft compared to the stroma, therefore the
stroma is modeled as a rigid substrate for simplicity, and the flow field vanishes at z = 0.
More general model that includes the viscoelastic properties of the stroma will be deferred to
future studies. The apical surface of the tissue is assumed to be free, and the free boundary
5
condition that includes contribution from surface tension of the tissue and external pressure
is applied.
The steady state velocity field of Eqs. (1)-(3) with these boundary conditions is v ∗
x =
0 kp(z ′)dz ′. The superscript ∗ is used to denote the steady state values.
v ∗
y = 0 and v ∗
z (z) =R z
Since v ∗
z should vanish at the apical surface, we have the following steady state condition
0 =Z H ∗
0
kp(z)dz,
(4)
where H ∗ is the thickness of the tissue in the steady state. Since kp is a decreasing function
of z, for the solution of Eq. (4) to exist, the tissue needs to have kp(0) > 0 and kp(H ∗) < 0.
Intuitively, this means that cell division events in the lower part of the tissue should be
balanced by cell apoptosis in the upper part of the tissue. We emphasize that our model
is independent of the details of kp(z), the only constraint for kp(z) is to have steady-state
tissue height H ∗ > 0 which satisfies Eq. (4).
A homeostasis state of a tissue is a steady state that is stable against small perturbations.
To check if the steady state in this model corresponds to a homeostasis state, we consider
a perturbation that takes the tissue height from H ∗ to H(x, t) = H ∗ + δH(x, t). Similarly,
vk = v ∗
k + δvk, p = p∗ + δp, etc. The linearized equations for the perturbed system are
and
∂lδvl = 0,
− ∂kδp + η∂l∂lδvk = 0.
(5)
(6)
Note that the perturbed tissue is translationally invariant in the y-direction.
For a perturbation with δH, δvi, δp ∼ eiqx+ωt. The no-slip boundary condition at z = 0
gives δvxz=0 = δvzz=0 = 0. The linearized boundary conditions at z = H ∗ are
− δpz=H ∗ + 2η∂zδvzz=H ∗ = σ∂x∂xδH,
and
η (∂zδvxz=H ∗ + ∂xδvzz=H ∗) = −2ηkp(H ∗)∂xδH,
(7)
(8)
where σ is the surface tension of the apical surface. The above two equations describe the
stress balance in the normal and tangential directions of the apical surface. There is also a
6
linearized kinematic boundary condition connecting the evolution of the tissue surface and
the flow field
∂tδH = δvzz=H ∗ + kp(H ∗)δH.
(9)
The dispersion relation ω(q) can be calculated from the linearized equations and boundary
conditions of the perturbed system, straightforward algebra gives
The first term
ω(q) = ωmech(q) + ωphy(q).
ωmech ≡
2qH ∗ − sinh 2qH ∗
2(qH ∗)2 + (1 + cosh 2qH ∗)
σq
2η
(10)
(11)
is the same as the dispersion relation for the surface of a low Reynolds number simple fluid.
In the limit of large qH ∗ this term approaches σq/2η [37]. As Fig. 1 (a) shows, ωmech is
negative for all q, this means that the surface tension of the apical surface tends to stabilize
the steady state. The second term is
ωphy ≡
1 + cosh(2qH ∗)
2(qH ∗)2 + 1 + cosh(2qH ∗)
kp(H ∗).
(12)
It is always negative since kp(H ∗) < 0, i.e., the apical region of the tissue is rich in terminal
differentiated cells which do not divide but undergo apoptosis. The meaning of this term is
intuitive: cell proliferation and cell apoptosis help to bring a perturbed tissue back to the
steady state. Note that for given kp(H ∗), ωphy has a minimum at wavenumber q = Ξ/H ∗,
where Ξ is the solution of cosh Ξ = Ξ sinh Ξ. Numerically one finds that Ξ ∼= 1.2 and
ωphy(q = Ξ/H ∗) ∼= 0.7kp(H ∗).
In this simple model, it is convenient to think that the relaxation of the tissue towards the
steady state is due to apical surface tension and cell proliferation. Contribution from surface
tension grows as the wavenumber q of the perturbation increases, approaching σq/2η in the
limit of large qH ∗. Contribution from cell proliferation has a minimum at qH ∗ ∼ O(1).
B. Second model: effect of viscosity heterogeneity
In a typical stratified epithelium, proliferative cells are located close to the basal re-
gion [16, 17]. Due to their small abundancy, typically the division rate of proliferative cells
is large compared to the apoptosis rate of differentiated cells [38]. Since stress relaxation
in a tissue strongly depends on the division and apoptosis of the cells [7], the viscosity in
7
regions rich in proliferative cells should be smaller than the viscosity in the other regions
of the tissue. Detailed discussion of the relation between cell division/apoptosis and tissue
viscosity is presented in Sec. III B and the Appendix.
To consider position-dependent viscosity we modify our model in Sec. II A by assuming
that all the proliferative cells reside in the region z < hS, and differentiated cells are located
at z > hS. This is clearly an over simplification, nevertheless it captures the basic tissue
composition in a stratified epithelium. The net cell proliferation rate inside the tissue is
simply kp = rS for z < hS and kp = −rD for z > hS. Here rS is the cell division rate and rD
is the cell apoptosis rate. From Eq. (4), the proportion of proliferative region in the steady
state is related to rS and rD by
hS
H ∗ =
r
1 + r
,
(13)
where r ≡ rD/rS. We emphasize that the control parameter in typical experiments is r.
For example, r can be increased by increasing the apoptosis rate of differentiated cells [39].
For given r and hS, we can use Eq. (13) to obtain the steady state tissue height H ∗. As a
reminder, note that in this model kp(H ∗) = −rD.
With the aforementioned simplification, in this model we choose η = ηS for z < hS and
η = ηD for z > hS. The geometry and boundary conditions for the linearized tissue evolution
equations are the same as those for our first model, except for additional velocity and stress
continuity conditions at z = hS. The steady state of our second model therefore becomes
v ∗
x = v ∗
y = 0, and v ∗
z = rSz for z < hS, v ∗
z = rShS − rDz for hS < z < H ∗. The continuity
of normal stress at hS leads to a discontinuity of tissue pressure in our model that should
be smoothed out in a real tissue because the boundary between proliferative region and
differentiated region is diffuse, not infinitely sharp.
On the other hand, in our second model the dynamics close to the steady state are
qualitatively different from that of the first model. Straightforward calculation for a small
perturbation of the form δH ∼ eiqx+ωt around the steady state of our second model gives
the relaxation rate ω. We find that ω can still be expressed as the sum of two terms ωmech
and ωphy, but these two terms are different from the counterparts of our first model,
ωmech = Kmech(qH ∗, r, ηrel)
σq
2ηD
,
ωphy = Kphy(qH ∗, r, ηrel) rD.
8
(14)
(15)
The explicit expression of the dimensionless functions Kmech and Kphy are pretty lengthy,
we present them in the Appendix. These functions are plotted in Fig. 2 as these figures
help us to get a better idea of how they behave as the parameters ηrel, qH ∗ and r change.
Besides their shapes, it is important to point out two issues. First, ωmech is the same as the
dispersion relation of the upper surface of a phase-separated binary low Reynolds number
fluid with lower fluid viscosity ηS, thickness hS, and upper fluid viscosity ηD, thickness
H ∗ − hS. Second, in the limit ηrel = 1, Eq. (14) and (15) become Eq. (11) and (12) as we
expected.
Figure 2 shows Kmech and Kphy for ηrel 6= 1. It can be seen that Kmech is always negative
(stabilizing), and its magnitude increases with r. This is because the thickness of the lower
(less viscous) region of the tissue increases and the damping is slowed down as r increases.
On the other hand, Kphy is a non-monotonic function of r, as this term is originated from
cell division/apoptosis, it cannot be simply explained by the viscous properties of the tissue.
We will discuss more about the behavior of Kphy in the next section when the flow field in
the tissue is examined.
For fixed r and ηrel, Kphy has a minimum at qH ∗ = Ξ(r, ηrel). Note that, unlike our first
model, here Ξ depends on r and ηrel. Figure 3(a) shows Ξ versus r for different choices of
ηrel. It is clear that in our second model the minimum of ωphy also occurs at qH ∗ ∼ O(1).
Figure 3(b) shows that when r and ηrel are large, it is possible for Kphy to become positive.
That is, it is possible for cell division/apoptosis to either slow down the relaxation toward
the steady state, or even drive the tissue to an instability. This will be discused in details
in Sec. III A.
In this section ηrel and r are treated as independent variables. We will discuss the relation
between ηrel and r in Sec. III B.
III. RESULTS AND DISCUSSIONS
A. Flow induced by cell apoptosis on the apical surface
An immediate interesting question of our hydrodynamic theory of epithelium tissue is,
besides fluidizing the tissue, how cell turnover inside the tissue affects tissue dynamics. This
is considered in both our models.
In the first model, we have assumed that the net cell
9
proliferation rate in a tissue is the only property that depends on the distance from the
basal membrane, all other properties of the tissue are homogeneous. On the other hand,
in the second model, regions with different net cell proliferation rate are assumed to have
different viscosity.
Interestingly, these two models predict very different tissue dynamics,
revealing nontrivial effects of cell turnover on tissue dynamics.
In our first model (Sec. II A), the tissue is homogeneous except for the proliferative
property. The linearized dynamics near the steady state shows that the steady state is
always stable under small perturbations. The relaxation rate towards the homeostasis state
contains two terms (Eq. (10)). The first term ωmech (Eq. (11)) is driven by the surface tension
of the apical surface, it is the same as the dispersion relation for the liquid-air surface of a
low Reynolds number simple liquid. It always stabilizes the tissue and this rate increases
with the wavenumber q (Fig. 1 (a)).
The second term ωphy (Eq. (12)) comes from cell turnover in the tissue. To gain further
insight on the effect of this term to the dynamics of the tissue, in Fig. 4 (a)(b) the field lines
of (δvx, δvy), the deviation of velocity field from the steady state in a tissue is plotted for a
tissue with zero apical surface tension σ for two different values of qH ∗. Since σ = 0, the
relaxation of the tissue is completely driven by the mechanism represented by ωphy. It can
be clearly seen that there is a flow induced by cell apoptosis at the apical surface. As shown
schematically in Fig. 4(c), cell apoptosis close to the apical surface gives an effective surface
flow. This is the mechanism that drives the flow field shown in Fig. 4(a) and (b). Figure 4(b)
shows that when qH ∗ is far from Ξ (Ξ ≈ 1.2), the flow is restricted to region close to the
apical surface. On the other hand, Fig. 4(a) shows that when qH ∗ is closer to Ξ, the flow
can penetrate deeper into the tissue. As q further decreases to qH ∗ ≪ 1 the apical surface
becomes nearly flat, and the induced horizontal flow (not shown) is again small. Therefore
the effect of induced flow is significant when qH ∗ ≈ Ξ. This is reflected in Sec. II A as the
minimum of the magnitude of ωphy at qH ∗ of order unity. The fact that ωphy has a minimum
at qH ∗ = Ξ also indicates that although intuitively cell turnover helps to stabilize the tissue,
flow induced by apoptosis at apical surface slows down the relaxation of the tissue. Indeed,
as can be seen from Fig. 4 (c), the induced flow tends to push cells from regions where the
tissue is thin to regions where the tissue is thick. Anyway, combining these two effects, in
this model ωphy is still negative for all q. That is, our first model predicts that cell turnover
helps to stabilize the homeostasis state, although the stabilizing effect is relatively weak
10
when qH ∗ ∼ 1. It is interesting to note that Fig. 4(a) and Fig. 4(b) also shows that vortices
appears in the flow field when qH ∗ > Ξ, and they are absent when qH ∗ < Ξ.
Fig. 5 shows the field lines of (δvx, δvz) predicted by our second model for a perturbed
tissue with zero apical surface tension. Again the field lines for qH ∗ > Ξ look very different
from the field lines for qH ∗ < Ξ. As qH ∗ increases from qH ∗ < Ξ to qH ∗ > Ξ, vortices appear
in the induced flow, this is similar to our first model (see Fig. 4 (b)). Let hc be the height of
the center of the vortices. When qH ∗ > Ξ, Fig. 5 shows that for hc > hS, Kphy decreases as
r increases; however, when hc < hS, Kphy increases as r increases. When hc = hS, Kphy has
a minimum (not shown in the figure). On the other hand, when qH ∗ < Ξ, Kphy increases
monotonically with r. This shows that the shape of Kphy in Fig. 2(c) is related to both the
flow induced by cell division/apoptosis and the thickness of the proliferative region.
A very important result of our second model is that the sign of Kphy can be positive, i.e.,
it is possible that the overall effect of cell division/apoptosis is to hinder the relaxation of
the tissue towards the steady state. This occurs when the following two conditions are both
satisfied. (i) The viscosity ηS in the region h < hS is sufficiently small compared to ηD (the
viscosity in the region z > hS). (ii) The relative thickness of the proliferative region hS/H ∗,
is sufficiently large. In Sec. III B we will discuss the relation between the distribution of
proliferative cells and viscosity, there we will show that indeed in general ηrel = ηD/ηS > 1.
However, a tissue that satisfies condition (ii) is likely not a normal tissue, as a healthy
mature epithelium tissue usually has a relatively thin layer of proliferative cells, a tissue
with large hS/H ∗ is more likely to be found during development or regeneration.
Stability diagrams of our second model are shown in Fig. 6 (a) and (b). It is convenient
to introduce a dimensionless parameter σq/(2ηDrD) to describe the relative importance of
the apical surface tension and cell turnover in tissue dynamics. As can be seen by comparing
Fig. 6 (a) and (b), tissues with small σq/(2ηDrD) are relatively easy to become unstable,
this is due to the relatively weak stabilizing effect from the apical surface tension.
Intuitively cell turnover does not necessarily drive a tissue unstable. Indeed we know from
our previous discussion that sometimes cell turnover helps to stabilize the homeostasis state,
while sometimes cell turnover drives a steady state unstable. Therefore it is interesting to see
when cell turnover stabilizes or destabilizes a steady state. Fig. 6 (c)-(f) shows how growth
rate of small perturbations close to the steady state changes with our model parameters.
It suggests that increasing rS and ηS helps to stabilize the steady state, as the peak of the
11
growth rate decreases; on the other hand, increasing rD and ηD helps to destabilize the
steady state, as the maximum growth rate of small perturbations increases. This further
confirms the intuition suggested by Fig. 6(a) and 6(b).
Now we can summarize the role of cell turnover in the dynamics of a tissue close to a
steady state. When the tissue height is perturbed from its steady state value, cell turnover
drives the tissue toward the steady state. On the hand, cell apoptosis on the apical surface
induces a flow in the tissue that brings more cells to thicker regions and pushes cells away
from regions with smaller thickness. For a normal tissue, proliferative cells occupy a small
region close to the basal surface, the overall effect of these mechanisms to tissue dynamics
always helps to stabilize the tissue. On the other hand, for a tissue with a large proportion
of proliferative cells, when the viscosity of the proliferative region of the tissue is sufficiently
small compare to the viscosity of the non-proliferative region, the induced flow can destabilize
the steady state, and the first instability occurs at wavelength 1/q ∼ H ∗.
B. Cell division, cell apoptosis, and tissue viscosity
Sec. II B stated that the viscosity in a tissue depends on the local tissue composition.
A derivation of the viscosity from a general hydrodynamic model of stratified epithelium
tissue is presented in the Appendix, where it is assumed that the stress relaxation process
has negligible contribution from the turnover of junction proteins, and cell division/apoptosis
produce force distribution that dominates the active stress inside a tissue. This approach
generalizes the earlier theory of Ranft et al. [7] by explicitly including the contribution from
different types of cells in a tissue [34, 36]. Our analysis shows that the viscosity η in a tissue
can be expressed as
η =
µσ0
ρh(rSdS − rD dD)ΛS + rD dDi
,
(16)
where ΛS = ρS/ρ, ρS is the number density of proliferative cells and ρ is the total cell
number density. µ is the elastic shear modulus of the tissue, rS is the cell division rate and
rD is the cell apoptosis rate, dS, dD are characteristic strengths of the force dipoles during
cell division and apoptosis, respectively. σ0 characterizes the response of cell polarity to the
stress in a tissue, cells in a tissue with small σ0 are more likely to be polarized by mechanical
stress. Equation (16) suggests that viscosity in a tissue depends on the local proportion of
12
proliferative cells. This supports the argument we made in Sec. II B.
Since in our model all the proliferative cells are localized in the basal (z < hS) region,
ΛS = 1 for z < hS and ΛS = 0 for z > hS. The viscosity in the tissue can be expressed as
µσ0
ρrS dS
= ηS, z < hS
µσ0
ρrD dD
= ηD, z > hS.
(17)
η(z) =
From Eq. (17), it is clear that the relative viscosity ηrel ≡ ηD/ηS and the relative proliferative
rate r ≡ rD/rS are not independent. In fact one can express ηrel as
ηrel =
η0
r
,
(18)
where η0 ≡ dS/ dD > 0 is a dimensionless parameter.
Equation (17) indicates that, for given η0, for a tissue to have large ηrel, the magnitude of
r needs to be small. On the other hand, Eq. (13) suggests that as r decreases, the proportion
of proliferative region in the tissue also decreases. Putting these two results together, Fig. 7
shows ηrel versus r for a few choices of η0. For the contribution from ωphy to drive the tissue
to an instability, the magnitude of η0 has to be large. Thus a tissue with small dD is more
likely to have a non-flat homeostasis state.
Note that the viscosity shown in Eq. (16) is derived for a tissue whose cell adhesion protein
turnover has negligible contribution to its dynamics. For a tissue in which the proliferative
cells have small number of intercellular adhesion sites compared to terminal differentiated
cells, ηrel could be greater than the magnitude predicted by Eq. (18). Therefore when both
turnover of cell adhesion proteins and active force dipoles due to cell division/cell apoptosis
contribute significantly to tissue dynamics, one still should treat r and ηrel as independent
parameters.
IV. CONCLUSION
Although biological tissue exhibit high plasticity that allows remodeling, the homeostasis
state is regulated such that its architecture is robust against intrinsic and external fluc-
tuations [2, 36].
In this article we have studied how local tissue composition affects its
viscous properties, and how viscous properties in turn affects the dynamics of a tissue close
to a steady state of a stratified epithelium. For a tissue with proliferative cells located
13
close to basal membrane, and differentiated cells located close to apical surface, the lower
region of the tissue has a different viscosity compared to the upper region. Since cell di-
vision/apoptosis inevitably induces a flow in this tissue, the position-dependent viscosity
in the tissue has a profound effect on its dynamics. Our analysis shows that usually cell
turnover is regulated to restore the steady state, but the flow induced by call apoptosis
close to the apical surface induces a flow that hinder this relaxation process. As a result,
perturbations with wavelengths comparable to the thickness of the tissue has the slowest
relaxation rate toward the steady state. When the tissue has a thick proliferative region,
and the viscosity of the proliferative region is significantly smaller than the other region, the
steady state can even become unstable.
Our model is quite general, thus our prediction should hold in general for stratified
continuous self-renewal tissues.
In the context of stratified epithelium, many undulating
patterns, for example rete peg [23 -- 25], occur at basal lamina, and its mechanical origin
has been explained by previous theoretical works [29, 33 -- 35]. On the other hand, our
result provides a possible route to trigger epithelial apical surface undulations. Although
an instability due to the induced flow described by our model is unlikely to happen in
a normal stratified tissue, such mechanism could be important during embryogenesis and
wound healing. Furthermore, when two different tissues are in contact [12], the viscosity
difference between these two tissues could also lead to interesting hydrodynamic instabilities.
These possibilities will be explored in our future works. It is also worth mentioning that
although in principle apical surface patterns can be a result of inhomogeneous distribution
of growth factors in the tissue [34], our model reveal that pure mechanical force is sufficient
to induce apical surface patterns.
To test our theory by in vitro experiments, one could try to increase the death rate
of differentiated cells to increase r, and decrease the tissue apical surface tension to make
the contribution from ωphy relatively more significant. In the case of mammalian olfactory
epithelium, a higher death rate can be induced by unilateral olfactory bulbectomy [39]. The
surface tension origins from the cell-cell adhesion [1, 2], and it could be decreased by, for
example the protease digestion procedure [40].
Several modifications can be made to make our model more general. For example, the
solid stroma in our model can be easily replaced by a soft stroma. Another important
improvement that we will make in the future work is to develop a model in whch cell
14
distribution is not pre-defined. It is important to check how homeostasis state and pattern
formation is achieved in a stratified epithelium with a more general hydrodynamic model.
The dynamics of cell lineage [34, 41, 42] will have to be taken into account in this future
work.
ACKNOWLEDGMENTS
The authors would like to thank support from the Ministry of Science and Technology,
Taiwan (grant number 102-2112-M-008-008-MY3) and NCTS. The authors are also grateful
for stimulating discussions with J-F Joanny (ESPCI), and helpful discussions with C-M
Chen (National Yang-Ming University, Taiwan).
Appendix A: Kphy and Kmech
1. Kphy and Kmech
The analytical form of Kphy and Kmech in Section 2.2 are
Kmech =
(ηrel + 1)2(2qH ∗ − sinh qH ∗) + (ηrel − 1)2µ1(qH ∗, hS)
(ηrel + 1)2(1 + 2(qH ∗)2 + cosh 2qH ∗) + (ηrel − 1)2µ2(qH ∗, hS)
,
Kphy =
(ηrel + 1)2[2(qH ∗)2] + (ηrel − 1)2µ3(qH ∗, hS)
(ηrel + 1)2[1 + 2(qH ∗)2 + cosh 2qH ∗] + (ηrel − 1)2µ2(qH ∗, hS)
− 1,
(A1)
where
µ1 =2qH ∗{1 − 2hS[1 + 2(qH ∗)2](hS − 1)hS} + 4qH ∗(hS − 1) cosh(2qH ∗hS) − sinh[2qH ∗(1 − 2hS)]
+ 2[1 + 2(qH ∗)2h2
S] sinh[2qH ∗(1 − hS)],
(A2)
µ2 =1 + 2(qH ∗)2{1 + 4(hS − 1)hS[1 + (qH ∗)2(hS − 1)hS]} + cosh[2qH ∗(1 − 2hS)]
− 2[1 + 2(qH ∗)2h2
S] cosh[2qH ∗(hS − 1)] − 2[1 + 2(qH ∗)2(hS − 1)2] cosh(2qH ∗hS),
(A3)
µ3 = 2(qH ∗)2{1 + 2hS[−2 + (1 + 2(qH ∗)2(hS − 1)2)hS]} + 2(hS − 1)2 cosh(2qH ∗hS). (A4)
Here hS ≡ hS/H ∗. Although these expressions are complicated, they can be obtained from
straightforward algebra.
15
Appendix B: viscosity in a tissue
In this Appendix we first construct a general model for continuous self-renewing tissues
that includes both spatial and cell lineage dynamics. By assuming that cell division and
apoptosis dominates the stress relaxation in a tissue, we show that a tissue behaves as
a viscoelastic material, and the viscosity in a tissue depends on local tissue composition.
The parameters in the theory can be estimated from existing experimental data of model
systems. The analysis in this Appendix follows closely the work of Ranft et al. [7]. The main
difference is that in our model the cell lineage, which has been identified as the fundamental
unit of tissue and organ development [36], is also taken into account.
There are proliferative cells and terminal differentiated (TD) cells in a tissue. Typically,
proliferative cells include stem cells and transit-amplifying (TA) cells differentiated from
stem cells. For simplicity we ignore this internal conversion of proliferative cells, only mark
that proliferative cells can undergo cell division, and the daughter cells have certain chance
to become terminal differentiated cells. On the other hand, TD cells do not divide, they
only undergo programmed cell death (apoptosis). Similar to other models [34, 36, 41], we
neglect apoptosis of proliferative cells.
Because of the coupling between the cortical network and the adhesion proteins, cells in
a tissue are tightly connected [2, 5], and diffusion of cells inside the tissue can be neglected
[44]. Denote ρS as the number density of proliferative cells, ρD as the number density of TD
cells. Taking into account cell division/apoptosis, and advection of cells by the flow inside
the tissue, the continuity equations for ρS and ρD can be written as
∂tρS + ∂l(vlρS) = rS(2pS − 1)ρS,
∂tρD + ∂l(vlρD) = 2rS(1 − pS)ρS − rDρD,
(B1)
(B2)
where Einstein summation convention is used to simplify the notation, rS and rD are the
division rate of proliferative cells and apoptosis rate of TD cells, respectively. pS is the
self-renewal probability of proliferative cells, and vi (i = x, y, z) is the i-th component of the
flow field.
The total cell density at position r is ρ(r, t) = ρS(r, t) + ρD(r, t). The fraction of the pro-
liferative cells at r is ΛS(r, t) = ρS(r, t)/ρ(r, t). From Eq. (B1)(B2), the evolution equations
16
for ρ and ΛS are
Dtρ = ρ[(rS + rD)ΛS − rD − ∂lvl],
DtΛS = [2rSpS − rS + rD − (rS + rD)ΛS]ΛS,
(B3)
(B4)
where Dt ≡ ∂t + vl∂l is the material derivative.
We assume that, in the absence of active processes such as cell division and cell apoptosis,
a tissue behaves as a linear elastic material, thus there is a linear relation between the change
of elastic stress ∆σE
ik and the change of elastic strain ∆uik, ∆σE
ik = (K − 2µ/3)∆ullδik +
2µ∆uik, where K is the compressional modulus, and µ is the shear modulus [45]. Besides
elastic stress, cell division and cell apoptosis also exert forces to the tissue [46]. Because
these processes exert no net force and torque on the tissue [47, 48], we model them as
symmetric force dipoles [7, 47, 53]. For a cell division/apoptosis event occurring at r = r0,
the change of stress is related to the active force dipole dik through ∆σA
ik = −dikδ(r − r0),
where δ(r − r0) is the Dirac delta function, and superscript "A" denotes stress created by
active cell division/apoptosis events. Since inertia is negligible [49], force balance condition
gives
∂iσik = 0,
(B5)
where σik ≡ σE
ik + σA
ik is the total stress in the tissue.
Due to cell division and apoptosis, a unique reference state for the strain does not exist.
However, the difference of strain between subsequent states still can be defined [7]. Let the
frame of reference flow and rotate with local flow and vortex, the evolution equation for the
stress can be written as
DJ
t σik = (K −
2
3
µ)vllδik + 2µvik + DJ
t σA
ik,
(B6)
where DJ
t σik ≡ Dtσik + Ωilσlk + Ωklσil is the Jaumann derivative. Here vik ≡ (∂ivk + ∂kvi)/2
is the strain rate tensor, and Ωik ≡ (∂ivk − ∂kvi)/2 is the vorticity tensor.
It is convenient to decompose the total stress tensor into isotropic and traceless parts, i.e.,
σik = −pδik + σik, where p ≡ −σll/3 is the tissue pressure, and σik is the traceless (deviator)
part of total stress tensor. Physically the isotropic part of the stress tensor is related to the
change of local tissue volume, and σik is related to the distortion of local tissue shape [45].
The same decomposition can also be applied to active stress tensor and strain rate tensor,
i.e. σA
ik = −pAδik + σA
ik and vik = (1/3)vllδik + vik. Now the stress evolution Eq. (B6) can be
17
expressed as
and
Dtp = −Kvll + DtpA,
DJ
t σik = 2µvik + DJ
t σA
ik.
(B7)
(B8)
To complete our model, we still need the constitutive equations for DtpA and DJ
t σA
ik.
Active stress is induced by cell division and cell apoptosis, we first consider the force dipole
created by cell division. Let unit vector li denote the direction of the cell division mitotic axis
(Fig. 8), follow [54], the force dipole produced by a cell division event can be approximated
by
ik = 3dSlilk.
ddiv
(B9)
Here dS characterizes the strength of the dipole. Typically dS > 0 because cell division
makes tissue grow. The magnitude of dS can be deduced from experiments. Since different
dividing cells may have different mitotic axis, the macroscopic properties is related to the
average of dik in a small region.
hddiv
ik i = 3dShlilki = dS(δik + qik),
(B10)
where qik ≡ h3lilk − δiki is the nematic order tensor in liquid crystal literatures. In general,
tissue growth and cell division/apoptosis processes are intrinsically anisotropic [5, 7, 50 -- 52],
and the orientation of the cell motitic axis depends on local tissue distortion [7, 55, 56]. To
linear order, this relation can be expressed as
qik − q0
ik =
σik
σ0
,
(B11)
where σ0 > 0 is a constant, and q0
q0
ll = 0. Denote the eigenvalues of qik − q0
ik as q(i), it is convenient to write
ik is the intrinsic anisotropic tensor, which should satisfy
q(i) = 3hcos2 θii − 1,
(B12)
where θi is the direction angle of mitotic axis relative to the i-th principle axis of qik − q0
ik.
From Eq. (B11) we have
Si =
1
6σ0 3σ(i) −
σ(k)! ,
3
Xk=1
18
(B13)
where Si ≡ (3hcos2 θii − 1)/2 is the three-dimensional order parameter relative to i-th
ik, and σ(i)'s are the i-th eigenvalue of σik. From this relation, one
principle axis of qik − q0
can estimate σ0 from experimental measurements [56].
Cell apoptosis, a key mechanism of tissue morphogenesis [57, 58], also exert force on the
surrounding environment (Fig. 8). During apoptosis, a cell rapidly develops an actomyosin
ring around its periphery and signal to neighboring cells to induce "purse-string"-like con-
tractility in the neighboring cells [59]. This "purse-string" contraction depends on myosin
activity, which can be elevated by applying mechanical forces to the tissues [60]. Therefore,
in general the force dipole exerted by cell apoptosis is anisotropic and dependent on the
local stress.
(B14)
hdapo
ik i = dDδik + dD(cid:18)q0
ik +
σik
σ0(cid:19) ,
the traceless part of this force dipole has an intrinsic contribution proportional to q0
ik and
a contribution induced by local stress. dD and dD are constants characterizing the strength
of the isotropic and the traceless part of the dipole. Typically dD < 0 since cell apoptosis
often leads to the decrease of tissue volume. On the other hand, there is no simple intuitive
argument to determine the the sign of dD.
Substitute Eq. (B10)(B14) into Eq. (B7)(B8) leads to
Dtp = −Kvll + ρ[(rSdS − rDdD)ΛS + rDdD],
(1 + τ DJ
t )σik = 2ηvik + σI
ik,
(B15)
(B16)
where σI
ik ≡ τ ρ[(rSdS − rD dD)ΛS + rD dD]q0
ik is the intrinsic active stress due to cell intrinsic
polarity, η ≡ τ µ, and the time scale τ that characterizes the active stress induced by cell
proliferation/apoptosis is given by
τ −1 ≡
ρ
σ0
[(rSdS − rD dD)ΛS + rD dD].
(B17)
From Eq. (B16) and Eq. (B17), we make the following remarks.
• For a tissue with no intrinsic cell polarity, σI
ik = 0. The stress evolution Eq. (B16)
reduces to the Maxwell model of a visco-elastic material with stress relaxation time τ
and viscosity η. At time scale greater than τ , the tissue behaves like a fluid [61],
σik = 2ηvik.
19
(B18)
It is important to note that the viscosity of the tissue depends on ρ and ΛS. This
supports the intuitive argument that we presented in the main text that the tissue
viscosity depends on the local density of proliferative cells and TD cells.
• From Eq. (B17) a sufficient condition for the tissue viscosity to be positive is dD > 0.
This is consistent with what has been observed in the experiments [8 -- 10].
• In general σI
ik is not zero. In this case, in the long time limit the stress of the tissue
satisfies
σik = 2ηvik + σI
ik.
(B19)
The additional term on the right-hand side of the above equation suggests that a
tissue in a flow-free state is in general under stress. This is the active stress from cells
with intrinsic polarity. It is important for tissue morphogenesis, for example, in the
imaginal disk of Drosophila [7, 52].
At time scale large compare to τ , the total stress tensor in a tissue reduces to
σik = −pδik + 2ηvik + σI
ik.
(B20)
Hence a tissue in this limit behaves as a viscous fluid coupled to a scalar field ΛS and an order
parameter which measures the intrinsic anisotropy of cell orientation. Moreover, Eq. (B20),
Eq. (B3-B5) and Eq. (B15) give a close set of equations describing tissue and cell lineage
dynamics. Note that the constitutive relation we used in the main text describes a tissue
with negligible intrinsic active stress.
By measuring traction stress exerted by the dividing Dictostelium cell on the flexible
poly-acrylamide gel, Tanimoto and Sano [54] found that the typical strength of force dipole
during cell division is about 1.2×10−13 N·m and the orientation of the force dipole coincides
with cell division axis [54]. This suggests that dS in Eq. (B9) can be measured. Although
Dictostelium does not form tissues, one can use this experimental measurement to get a
rough idea of the order of magnitude of dS as dS ∼ 4 × 10−14 N · m.
Experiments similar to [56] can be applied to measure σ0. In this experiment, Marel et al.
measured the average cell division orientation as a function of local strain rate in monolayer
Madin-Darby canine kidney (MDCK) cells and concluded that
2hcos2 φi − 1 ∼=
5
8
(λ1 − λ2),
20
(B21)
where λ1 and λ2 (λ1 > λ2) are the eigenvalues of strain rate tensor vik, and φ is the angle
between cell division axis and the principle axis of the eigenvalue λ1. By assuming the
monolayer tissue behaves as a Newtonian fluid, we can use the two-dimensional form of
Eq. (B13) ,
2hcos2 φi − 1 ∼=
¯η
¯σ0
(λ1 − λ2),
(B22)
where ¯η is the 2-dimensional viscosity. Using typical tissue viscosity η ∼= 105 Pa · s [10] and
MDCK monolayer height h ∼= 2 µm [62], a comparison between Eq. (B21) and Eq. (B22)
gives ¯σ0
∼= 5 Pa · m. Thus σ0 = ¯σ0/h ∼= 2.5 × 106 Pa.
For amnioserosa tissue, the magnitude of dD can be inferred from the apoptosis rate
measured from dorsal closure process of Drosopila. The measurement made by Toyama et
∼= 5.5 × 10−5 s−1 [57, 65]. Since during dorsal closure, amnioserosa cell only
al. gives rD
undergo cell apoptosis, Eq. (B17) is reduced to τ −1 = ρrD dD/σ0. Use tissue height h ∼= 5µm
[63], cell density ρ ∼= 2.2 × 10−3 µm−3 [57], and the relaxation time τ ∼= 100 s [64]. If we
∼= 2.1 × 10−7 N · m for
use the magnitude of σ0 measured from MDCK cells, we find dD
amnioserosa tissue. In principle, combining the measurement by [57] with an experiment
like [56] for amnioserosa tissue give us a better estimate for its dD.
In summary, in this Appendix we have presented a theoretical model to relate the viscosity
of a tissue to a few measurable parameters. The key parameters dS, dD, and σ0 of a tissue can
be measured from different experiments. Therefore it is possible to infer the hydrodynamic
properties of a tissue from existing experimental techniques.
[1] Gonzalez-Rodriguez D, Guevorkian K, Douezan S and Brochard-Wyart F 2012 Soft matter
models of developing tissues and tumors Science 338 910 -- 917
[2] Lecuit T and Lenne P F 2007 Cell surface mechanics and the control of cell shape, tissue
patterns and morphogenesis Nat. Rev. Mol. Cell Biol. 8 633 -- 644
[3] Marchetti M C, Joanny J F, Ramaswamy S, Liverpool T B, Prost J, Rao M and Simha R A
2013 Hydrodynamics of soft active matter Rev. Mod. Phys. 85 1143 -- 1198
[4] Yang S, Khare K and Lin P C 2010 Harnessing surface wrinkle patterns in soft matter Adv.
Funct. Mater. 20 2550 -- 2564
21
[5] Perez-Moreno M, Jamora C and Fuchs E 2003 Sticky business: Orchestrating cellular signals
at adherens junctions Cell 4 535 -- 548
[6] Thoumine O, Lambert M, Mege R M and Choquet D 2006 Regulation of n-cadherin dynamics
at neuronal contacts by ligand binding and cytoskeletal coupling Mol. Biol. Cell 17 862 -- 875
[7] Ranft J, Basan M, Elgeti J, Joanny J-F, Prost J and Julicher F 2010 Fluidization of tissues
by cell division and apoptosis Proc. Nat. Acad. Sci. USA 107 20863 -- 20868
[8] Bonnet I, Marcq P, Bosveld F, Felter L, Bellaiche Y and Graner F 2012 Mechanical state,
material properties and continuous description of an epithelial tissue J. R. Soc. Int. 9 2614 --
2623
[9] Foty R A, Forgacs G, Pfleger C and Steinberg M S 1994 Liquid properties of embryonic tissues:
Measurement of interfacial tensions Phys. Rev. Lett. 72 2298 -- 2301
[10] Forgacs G, Foty R A, Shafrir Y and Steinberg M S 1998 Viscoelastic properties of living
embryonic tissues: A quantitative study Biophys. J. 74 2227 -- 2234
[11] Thoumine O and Ott A 1997 Time scale dependent viscoelastic and contractile regimes in
fibrobalsts probed by microplate manipulation J. Cell Sci. 110 2109 -- 2116
[12] Basan M, Risler T, Joanny J-F, Sastre-Garau X and Prost J 2009 Homeostatic competition
drives tumor growth and metastasis nucleation HFSP J. 3 265 -- 272
[13] Rizvi A Z and Wong M H 2005 Epithelial stem cells and their niche: There is no place like
home Stem Cells 23 150 -- 165
[14] Reece J B, Urry L A, Cain M L, Wasserman S A, Minorsky P V and Jackson R B 2010
Campbell Biology 9th Edition (Benjamin Cummings)
[15] Genzer J and Groenewold J 2006 Soft matter with hard skin: From skin wrinkles to templating
and material characterization Soft Matter 2 310 -- 323
[16] Watt F M and Hogan B L M 2000 Out of eden: Stem cells and their niches Science 287
1427 -- 1430
[17] Moore K A and Lemischka I R 2006 Stem cell and their niches Science 311 1880 -- 1885
[18] Chowdhury F, Na S, Li D, Poh Y-C, Tanaka T S, Wang F and Wang N 2010 Material
properties of the cell dictate stress-induced spreading and differentiation in embryonic stem
cells Nat. Mater. 9 82 -- 88
[19] Babahosseini H, Ketene A N, Schmelz E M, Robert P C and Agah M 2014 Biomehcanical
profile of cancer stem-like/tumor-initiating cells derived from a progressive ovarian cancer
22
model Nanomedicine: NCM 10 1013 -- 1019
[20] Li B, Cao Y-P, Feng X-Q, and Gao H 2012 Mechanics of morphological instabilities and surface
wrinkling in soft materials: a review Soft Matter 8 5728
[21] Cerda E and Mahadevan L 2003 Geometry and physics of wrinkling Phys. Rev. Lett. 90 074302
[22] Flynn C O and McCormack B A O 2009 A three-layer model of skin and its application in
simulating wrinkling Comput. Methods Biomech. Biomed. Eng. 12 125 -- 134
[23] Hale A R 1952 Morphogenesis of volar skin in human fetus Am. J. Anat. 91 147 -- 180
[24] Babler W J 1991 Embryologic development of epithelial ridges and their configurations Birth
Defects Orig. 27 95 -- 122
[25] Lavker R M and Sun T T 1983 Epidermal stem cells Adv. Dermatol. 21(S1) 121 -- 127
[26] Li B, Cao Y-P, Feng X-Q and Gao H 2011 Surface wrinkling of mucosa induced by volumetric
growth: Theory, simulation and experiment J. Mech. Phys. Solids 59 758 -- 774
[27] Li B, Cao Y-P, Feng X-Q, and Yu S-W 2011 Mucosal wrinkling in animal antra induced by
volumetric growth Appl. Phys. Lett. 98 153701
[28] Klein-Szanto A J P and Schroeder H E 1975 Architecture and density of the connective tissue
papillae of the human oral mucosa J. Anat. 123 93 -- 109
[29] Ciarletta P and Amar M Ben 2012 Papillary networks in the dermal-epidermal junctions of
skin: A biomechanical model Mech. Res. Commun 42 68 -- 76
[30] Kuchen M and Newell A C 2005 Fingerprint formation J. Ther. Biol. 235 71 -- 83
[31] Hannezo E, Prost J and Joanny J-F 2011 Instabilities of monolayered epithelia: Shape and
structure of villi and crypts Phys. Rev. Lett. 107 078104
[32] Ciarletta P 2013 Buckling instability in growing tumor spheroids Phys. Rev. Lett. 110 158102
[33] Basan M, Joanny J-F, Prost J and Risler T 2011 Undulation instability of epithelial tissues
Phys. Rev. Lett. 106 158101
[34] Ovadia J and Nie Q 2013 Stem cell niche structure as an inherent cause of undulating epithelial
morphogenesis Biophy. J. 104 237 -- 246
[35] Risler T and Basan M 2013 Morphological instabilities of stratified epithelia: A mechanical
instability in tumor formation New J. Phys. 15 065011
[36] Lander A D, Gokoffski K K, Wan F Y-M, Nie Q and Calof A L 2009 Cell lineages and the
logic of proliferative control PLoS Biol. 7 0084 -- 0100
[37] Wu E S and Webb W W 1973 Critical liquid-vapor interface in SF6. II. thermal excitations,
23
surface tension, and viscosity Phy. Rev. A 8 2077 -- 2084
[38] Wu H H, Ivkovic S, Murray R C, Jaramillo S, Lyons K M, Johnson J E and Calof A L 2003
Autoregulation of neurogenesis by gdf11 Neuron 37 197 -- 207
[39] Holcomb J D, Mumm J S and Calof A L 1995 Apoptosis in the neuronal lineage of the mouse
olfactory epithelium - regulation in-vivo and in-vitro Dev. Biol. 172 307 -- 323
[40] Weinberg R A 2007 The Biology of Cancer (Garland Science)
[41] Chou C-S, Lo W-C, Gokoffski K K, Zhang Y-T, Wan F Y-M, Lander A D, Calof A L and
Nie Q 2010 Spatial dynamics of multistage cell lineages in tissue stratification Biophy. J. 99
3145 -- 3154
[42] Hannezo E, Prost J and Joanny J-F 2014 Growth, homeostatic regulation and stem cell
dynamics in tissue J. R. Soc. Interface 11 20130895
[43] Potten C S 1974 The Epidermal Proliferative Unit: The Possible Role of the Central Basal
Cell Cell Tissue Kinet 7 77 -- 88
[44] Friedl P and Gilmour D 2009 Collective cell migration in morphogenesis, regeneration and
cancer Nat. Rev. 10 445 -- 457
[45] Landau L D, Pitaevskii L P and Lifschitz E M 1986 Theory of Elasticity (3rd edition) (Elservier
Butterworth-Heinemann)
[46] Kim J H, Dooling L J and Asthagiri A R 2010 Intercellular mechanotransduction during
multicellular morphodynamics J. R. Soc. Interface 7 S341 -- S350
[47] Schwarz U S and Safran S A 2002 Elastic interactions of cells Phys. Rev. Lett. 88 048102
[48] Schwarz U S, Balaban N Q, Riveline D, Bershadsky A, Geiger B and Safran S A 2002 Calcu-
lation of forces at focal adhesions from elastic substrate data: the effect of localized force and
the need for regularization Biophys. J. 83 1380 -- 1394
[49] Rauzi M and Lenne P F 2011 Cortical forces in cell shape changes and tissue morphogenesis
Curr. Top Dev. Biol. 95 93 -- 144
[50] Rauzi M, Verant P, Lecuit T and Lenne P F 2008 Nature and anisotropy of cortical forces
orienting drosopila tissue morphogenesis Nat. Cell Biol. 10 1401 -- 1410
[51] Bryant D M and Mostov K E 2008 From cells to organs: Building polarized tissue Nat. Rev.
Mol. Biol. 9 887 -- 901
[52] Bittig T, Wartlick O, Kicheva A, Gonzalez-Gaitan M and Julicher F 2008 Dynamics of
anisotropic tissue growth New J. Phys. 10 063001
24
[53] Bischofs I B, Safran S A and Schwarz U S 2004 Elastic interactions of active cells with soft
materials Phys. Rev. E 69 021911
[54] Tanimoto H and Sano M 2012 Dynamics of traction stress field during cell division Phys. Rev.
Lett. 109 248110
[55] Thery M, Jimenez-Dalmaroni A, Racine V, Bornens M and Julicher F 2007 Experimental and
theoretical study of mitotic spindle orientation Nature 447 493 -- 497
[56] Marel A-K, Podewitz N, Zorn M, Radler J and Elgeti J 2014 Alignment of cell division axes
in directed epithelial cell migration New J. Phys. 16 115005
[57] Toyama Y, Peralta X G, Wells A R, Kiehart D P and Edwards G S 2008 Apoptotic force and
tissue dynamics during drosopila embryogenesis Science 321 1683 -- 1686
[58] Monier B, Gettings M, Gay G, Mangeat T, Schott S, Guarner A and Suzanne M 2015 Apico-
basal forces exerted by apoptotic cells drive epithelium folding Nature 518 245 -- 248
[59] Rosenblatt J, Raff M C and Cramer L P 2001 An epithelial cell destined for apoptosis signals
its neighbors to extrude it by an actin- and myosin-dependent mechanism Curr. Biol. 11
1847 -- 1857
[60] Fernandez-Gonzalez R, de Matos Simoes S, Roper J C, Eaton S, and Zallen J A 2009 Myosin
ii dynamics are regulated by tension in intercalating cells J. A. Dev. Cell 17 736 -- 743
[61] Mase G T, Smelser R E and Mase G E 2009 Continuum Mechanics for Engineers (3rd edition)
(CRC Press)
[62] Hoh J H and Schoenenberger C-A 1994 Surface morphology and mechanical properties of
mdck monolayers by atomic force microscopy J. Cell Sci. 107 1105 -- 1114
[63] Saias L, Swoger J, D'Angelo A, Hayes P, Colombelli J, Sharpe J, Salbreus G and Solon J 2015
Decrease in cell volume generates contractile forces driving dorsal closure Dev. Cell 33 1 -- 11
[64] Solon J, Kaya-Copur A, Colombelli J and Brunner D 2009 Pulsed forces timed by a ratchet-like
mechanism drive directed tissue movement during dorsal closure Cell 137 1331 -- 1342
[65] Note that the apoptosis rate measured in [57] corresponds to Ntotal times rD in our model.
Ntotal is the number of amnioserosa cells just before dorsal closure. Typically Ntotal ≈ 100.
25
FIG. 1. 2ηωmech/σq (blue curve) and ωphy/kp(H ∗) (red curve). ωmech(q) decreases monotonically
towards σq/2η as qH ∗ ≫ 1, ωphy has a minimum at qH ∗ ∼ 1.
26
FIG. 2.
(a) Kmech as a function of r and qH ∗ for ηrel = 5. (b) Kmech versus r for qH ∗ = 2,
ηrel = 5 (solid blue curve), qH ∗ = 0.6, ηrel = 5 (solid red curve), qH ∗ = 2, ηrel = 4 (dashed
blue curve), and qH ∗ = 0.6, ηrel = 4 (dashed red curve). (c) Kphy as a function of r and qH ∗ for
ηrel = 5. (d) Kphy/rD versus r for qH ∗ = 2, ηrel = 5 (solid blue curve), qH ∗ = 0.6, ηrel = 5 (solid
red curve), qH ∗ = 2, ηrel = 4 (dashed blue curve), and qH ∗ = 0.6, ηrel = 4 (dashed red curve).
27
FIG. 3. (a) Ξ as a function of r for ηrel = 0.4 (blue), 1 (red), and 5 (orange). (b) This plot shows
the least negative value of Kphy for given r, ηrel, i.e., Kphy(qH ∗ = Ξ, r, ηrel). The dotted curve
shows where Kphy(qH ∗ = Ξ, r, ηrel) is zero.
28
FIG. 4.
(a) and (b): Field lines of (δvx, δvy) calculated from our first model for a perturbed tissue
with tissue apical surface tension σ = 0. It can be seen that the topology of the flow field depends
on the magnitude of qH ∗. qH ∗ = 0.8 in (a), qH ∗ = 2 in (b). The red arrows are the stream lines,
the dashed black lines are the apical surface of the perturbed tissue, and the background gray-level
indicates the magnitude of δv. To generate the flow field, we have chosen kp(H ∗) = −1, H ∗ = 2,
and δH = 0.01 cos(qx). (c) Cell death at apical surface induces a flow field that is indicated by
the empty arrows. Close to the apical surface this flow field has a horizontal component indicated
by the blue arrows. The solid curve is the apical surface at the present moment, and the dashed
curve indicates the apical surface at a later time.
29
FIG. 5. Field lines of (δvx, δvy) in the second model for σ = 0, and Kphy versus r for qH ∗ = 0.8
(< Ξ for all r) and qH ∗ = 2.0 (> Ξ for all r). For qH ∗ = 0.8, Kphy increases monotonically with
r. For qH ∗ = 2, there are vortices in the induced flow. Kphy decreases with r when hc > hS, Kphy
increases with r when hc < hS . hS is indicated by the green lines, and hc is the z-position of the
vortex center. To generate the flow fields, we have chosen ηrel = 5, kp(H ∗) = −1, H ∗ = 2, and
δH = 0.01 cos(qx).
30
FIG. 6.
(a)(b) Gray scale indicates th relaxation rate of a perturbed tissue scaled by rD for
qH ∗ = Ξ. (a) σq/(2ηDrD) = 0.4, and (b) σq/(2ηDrD) = 0.05. (c)-(f) relaxation rate (in units
of 1/day) of a tissue as a function of the wavelength of perturbation. Solid curves: rS = 1 d−1,
rD = 30 d−1, ηS = 10 MPa s, ηD = 50 MPa s, σ = 1 m Nm−1, and hS = 30 µm (the corresponding
magnitude of H ∗ is 31 µm). The dashed lines give the relaxation rate versus λ for the same
parameters except rS = 2 d−1 in (c), rD = 43 d−1 in (d), ηS = 20MPa s in (e), and ηD = 80MPa s
in (f).
31
FIG. 7. r versus ηrel for η0 = 30 (blue curve), 90 (red curve), and 150 (yellow curve). The short-
dashed curve is the stability boundary for a tissue with σq/(2ηDrD) = 0.2 and qH ∗ = Ξ. The
stability boundary is determined from the zero of the relaxation rate for perturbations with the
most dangerous wavelength.
32
FIG. 8. Cell division and apoptosis as a source of localized force inside tissue. (a) Cell division
can produce equal but opposite forces along the mitotic axis (dashed arrow). (b) Cell apoptosis
induce a purse-string (dashed line) which shrinks the cell surface and prevents gap formation. Both
processes create a force dipole acting on the surrounding environment (blue cells).
33
|
1205.6126 | 2 | 1205 | 2012-09-24T16:00:20 | Dynamical model for the full stretching curve of DNA | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech",
"q-bio.BM"
] | We present a phenomenological dynamical model able to describe the stretching features of the curve of DNA length vs applied force. As concerns the chain, the model is based on the discrete wormlike chain model with elastic modifications, which properly describes the elongation features at low and intermediate forces. The dynamics is developed under a double-well potential with a linear term, which, at high forces, accounts for the narrow transition present in the DNA elongation (overstretching). A quite good agreement between simulation and experiment is obtained. | physics.bio-ph | physics |
Dynamical model for the full stretching curve of DNA
Alessandro Fiasconaro1, 2, ∗ and Fernando Falo1, 3
1Departamento de F´ısica de la Materia Condensada,
Universidad de Zaragoza, 50009 Zaragoza, Spain
2Instituto de Ciencia de Materiales de Arag´on, CSIC -- Universidad de Zaragoza, 50009 Zaragoza, Spain
3Instituto de Biocomputaci´on y F´ısica de Sistemas Complejos, Universidad de Zaragoza, Zaragoza, Spain
(Dated: October 15, 2018)
We present a phenomenological dynamical model able to describe the stretching features of a
length vs applied force DNA curve. As concerning the chain, the model grounds on the discrete
worm-like chain model with the elastic modifications, which properly describes the elongation fea-
tures at low and intermediate forces. The dynamics is developed under a double well potential with a
linear term, which, at high forces, accounts for the narrow transition present in the DNA elongation
(overstretching). A quite good agreement between simulation and experiment is obtained.
PACS numbers: 87.15.-v, 36.20.-r, 87.18.Tt, 83.10.Rs, 05.40.-a
Introduction. The stretching curve of DNA has been
the subject of many studies, both theoretical and ex-
perimental. The first experimental study of the DNA
stretching was performed by Bustamante and co-workers
[1], where the DNA length was measured as a function of
the force applied to the chain. From the theoretical point
of view, the small and intermediate force range of the
DNA stretching curve can be understood and depicted
by means of the wormlike chain (WLC) model [2], where
the DNA is modeled as a flexible beam. For simulation
purposes, the WLC model can be discretized as a chain
of beads connected by sticks with the presence of an elas-
tic bending (discrete WLC) and improves the more naive
freely joined chain model [3]). This WLC model (and the
discrete version) fits with excellent precision the experi-
mental curve, making use of the correction introduced by
Odijk [4] that takes into account the extensibility of the
chain when the sticks are replaced by harmonic springs.
The most intriguing feature of the stretching of DNA
is presented in a subsequent experiment of Bustamante
and co-workers [5, 6], where a sudden elongation of the
chain is registered when a large force, around 70pN , is
applied.
The presence of this tension-induced overstretching
transition reveals the existence of two structurally differ-
ent DNA states. The first state, at low applied forces, is
the so called B-DNA conformation, where the base pairs
are helicoidally packed in their native state. The other
state, present at high forces, was first called the S state,
where the base pair distance is higher by about 70% than
that in the state B. The nature of the overstretched DNA
state remains controversial, and it seems that many ef-
fects could be responsible for the sudden DNA elongation
at large forces.
Molecular dynamics simulations revealed that the
DNA transition could correspond to a change of the he-
licoidal DNA shape to another one with a shorter helix
∗Electronic address: [email protected]
radius and a large tilt of the base pairs [6]. This result
seems also be confirmed by early spectroscopic studies
[7].
On the other hand, the suggestion that the enlarge-
ment is accompanied by a rotation of the bases giving
rise to a ladderlike structure in the S state of DNA, has
also been made [8]. Motivated by the experimental data
of Strick et al. [9], a possible linear coupling of the twist-
stretch variable was proposed [10, 11], and in this last
paper an estimation of the DNA twist stiffness was also
performed. More accurate measurements suggest that a
not completely ladderlike form of the DNA occurs in the
elongation from the B- to the overstretched DNA state.
In fact a remanent helicity of the DNA persists in its over-
stretched state of about one turn every 37.5 base pairs,
almost four times smaller than the natural helicity of one
turn per 10.5 base pairs [12]. Moreover, by avoiding the
writhing of the DNA double helix, a broadening in the
overstretching transition is observed.
Various statistical-mechanics models have been imple-
mented in order to mathematically describe the narrow
overstretching transition, which take into account the
WLC model [12 -- 14], with an increasing number of pa-
rameters in order to take into account the twist con-
straints [11, 15].
The existence of the S-DNA state as a hybridized state
of DNA has been the object of deep and still controver-
sial investigation, mainly arising because of the presence
of an asymmetric hysteresis in the pulling and relax-
ing DNA elongation curve, which appeared to indicate
a force-induced DNA melting [16 -- 18]. Further investiga-
tions, both numerical and experimental, were published
up until very recently, some of which confirmed [19 -- 22],
and some did not confirm [23 -- 25] this hypothesis.
In this work, we present a dynamical Langevin model
where a polymer is pulled under an asymmetric free en-
ergy potential, with the purpose of describing the narrow
DNA overstretched transition at high forces in a Landau-
Ginzburg landscape. This model proposes a simple phe-
nomenological way to describe the DNA dynamics, and
its results are in good agreement with the most relevant
FIG. 1: Scheme of linear chain in 3D space. In (a) the chain
is free to move in the thermal bath, while in (b) is pulled from
each end with a force Fp.
Fp = 0.0 pN
Fp = 31.0 pN
Fp = 68.5 pN
Fp = 92.8 pN
Fp = 123.8 pN
V(di)
120
100
80
60
40
20
0
-20
-40
-60
1
1.1
1.2
1.3
1.4
di
1.5
1.6
1.7
1.8
FIG. 2: Potential for different values of the force Fp acting
on the chain. The forces increase from the upper to the lower
curves.
experimental outcomes. The model we present doesn't
enter in the microscopic details of the DNA elongated
state, if melting is or is not responsible for it. The de-
scription presented is valid regardless of the specific mi-
croscopic state of the DNA. In this sense the model is
general and can be included as an ingredient into other
more complex processes like DNA translocation driven
by molecular and nanotechnological motors.
The chain model. We model the DNA molecule by a
three-dimensional (3D) polymeric chain of N dimension-
less monomers connected by nonharmonic springs (1). In
spite of the usual harmonic interaction [26], our elastic
potential energy is given by
Vel(di) =
ke
2
N
X
i=1
(di − l0)2(di − l1)2 + kldi,
(1)
where ke is the elastic parameter, ri is the position of the
i-th particle, di = di = ri+1 − ri is the distance be-
tween the monomers i+1 and i, and l0 and l1 are minima
of the potential representing approximately the equilib-
rium distance between adjacent monomers, at weak and
strong forces respectively. The potential is plotted in
Fig. 2. The parameter kl is chosen in such a way that the
middle of the force transition corresponds to equal prob-
abilities to cross the elongation potential barrier from
either left to right (enlargement) or right to left (con-
striction).
The model takes into account the bending energy with
2
I)
II)
IV)
1
0.8
dx
0.6
0.4
III)
Region I)
1
2
3
Fp
4
WLC-E
dx(F)
EXPT.
40
50
Fp (pN)
60
70
80
90
c
d
/
x
d
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0
10
20
30
FIG. 3: DNA length vs Force. Parameters: ts = 5 × 106,
N = 100, kB T = 4.1, kb = 10, ke = 1190, and kl = 89 (units
given in the text).
a term given by
Vben(θi) =
kb
2
N
X
i=1
[1 − cos(θi − θ0)],
(2)
where kb is the bending elastic constant, and θi is the
angle between the link di+1 and the link di.
The dynamics of the chain is given by the overdamped
equation of motion
ri = −∇iVel(di) − ∇iVben(θi) − p2kBT ξ(t),
(3)
where ξ(t) represents the thermal contribution as a Gaus-
sian uncorrelated noise, and the time t is scaled with the
damping γ as t → t/γ. Time units are then given in
sγ−1. The nabla operator is defined as ∇i = ∂/∂xii +
∂/∂yij + ∂/∂zik. Two forces Fp and −Fp act on the
chain on the first and last monomer respectively in order
to stretch the polymer. The figures show the length of
the DNA polymer as a function of the applied force Fp.
Results. We performed various computer simulations
with a Runge-Kutta stochastic algorithm for a long fixed
simulation time ts = 5 × 106 time units (t.u.), and
dt = 0.01. We set the units of energy and length to
be Eu = 4.1 pN nm (thermal energy at room tem-
perature) and lu = 5.3 nm, respectively. With these
units the parameters used in simulations are kBT = Eu,
ke = 1190Eu/l4
u, l0 = lu, l1 = 1.7lu, kb = 10Eulu,
and θ0 = 0. Thus the persistence length of the chain
is ξ = kb/kBT = 10lu = 53nm, which gives the correct
WLC model fit from the experimental data. The param-
eter ke is determined from the extended WLC model in
order to fit the DNA elongation data in region II, and
analogously for the bending parameter kb.
After a thermalization time of ts = 1 × 106 t.u., the
average of the lengths at all subsequent times has been
evaluated with respect to the contour length dc of the
polymer, with dc = (N − 1)l0, and the values are plot-
ted in Fig. 3, where a very good agreement between
the computer simulation (brown ⊡) and the experiments
(black ×) is shown [5].
The inset of the figure shows good agreement also at
low force values. The full line represents the prediction
of the WLC model in region I and II, which follows very
well both the simulations points and the experimental
ones. Figure 4 shows the density distribution of the link
length close to the transition.
It is there evident that
a gradual increase of the number of enlarged links (i.e.,
di ≈ 1.7) is produced by increasing the applied pull force
Fp.
The transition between the regimes II (B-DNA) and IV
(overstertched) arises in the very short force range. We
model this feature using the Ginzburg-Landau (GL) ef-
fective potential given in Eq. (1). This approach has also
been used to describe the B-DNA to Z-DNA transition
[27]. In the GL phase transition theory, an asymmetrical
contribution to the free energy gives rise to a discontin-
uous (first-order) phase transition with the temperature
T as control parameter [28]. Analogously, in the present
case, we obtain a sharp transition with the DNA elonga-
tion dx as order parameter, and the force Fp as control,
with the difference with respect to the standard behav-
ior that the intensity of thermal fluctuations is here an
external parameter, whose presence causes a relatively
smooth transition between the two states.
A dynamical fingerprint of discontinuous transitions
is the appearance of hysteretic behavior due to a slow
convergence to equilibrium states. This behavior has
been clearly observed in out-of-equilibrium experiments
on the overstretching transition [23, 24, 29]. The hys-
teresis curves are consequence of force ramps applied on
a time scale shorter than the DNA relaxation time. Using
waiting times of the order of 1 min the equilibrium state
is recovered and no hysteresis is observed [12]. Figure 5
shows how our model mimics this behavior. In fact, the
left panel of the figure presents a typical hysteresis curve
obtained by decreasing the simulation time, i.e., not al-
lowing the system to reach equilibrium. For a simulation
Fp = 65 pN
Fp = 67 pN
Fp = 69 pN
Fp = 71 pN
12
10
8
P(di)
6
4
2
0
0.9
1
1.1
1.2
1.3
1.4
di
1.5
1.6
1.7
1.8
1.9
FIG. 4: Probability density of the length of the links di for
different applied forces along the overstretching transition.
3
1.7
1.6
1.5
c
d
/
x
d
1.4
1.3
1.2
1.1
1
kl = 82
kl = 89
kl = 96
EXPT.
1.7
1.6
1.5
c
d
/
x
d
1.4
1.3
1.2
1.1
3×105
106
5×106
EXPT.
60
65
70
Fp (pN)
75
80
1
50 55 60 65 70 75 80
Fp (pN)
FIG. 5: Left panel: transition region with different applica-
tion times of the force, ts = 3 × 105, and ts = 106, shorter
than the relaxation time of the chain. Right panel: the same
for different values of the potential parameter kl, kl increasing
from left to right curves.
time of ts = 106 t.u., we already have a different path
for increasing forces than for decreasing ones (indicated
in the figures by the black arrows). For a shorter simula-
tion time ts = 3 × 105 t.u., the hysteresis area is bigger,
while for ts = 5 × 106 t.u. (squares in the figure), no
hysteresis is observed. This behavior expresses a typical
dynamical property of the model not visible in a static
equilibrium analysis.
Finally, we discuss the effect of the kl parameter on
the transition. The right panel of Fig. 5 shows how this
parameter is responsible in the model for the force value
where the transition occurs. This value can be related
to those parameters whose magnitude affects the over-
stretching transition: the value of pH [16] or the salt
concentration [17, 30].
Discussion. The picture of the DNA overstretching
transition is getting richer and richer as the number of
experiments increases. However, as we stated in the In-
troduction, the microscopic nature of the transition is
still controversial. An interesting approach to the prob-
lem is to study single-stranded DNA [31]. Although the
elongation features in that case could give some insight
in explaining the rich elastic properties of DNA, the pres-
ence of two plateaus in single-stranded DNA found in [31]
shows that the elasticity of DNA is much more complex
than previously thought.
It seems difficult to put to-
gether the features of a double-stranded DNA with those
of single-stranded DNA, where the force-elongation char-
acteristics depend on the nature of the specific piece of
DNA object of elongation.
Possibly related to this, other experimental works sug-
gest that DNA melting could be responsible for the
DNA overstretching features [16 -- 18, 21, 22]. Although
strong evidence in that direction appears reasonable,
many doubt still remain in this respect [23 -- 25, 32].
The model presented here does not consider any mi-
croscopic aspects of the transition but rather adopts, in
the GL spirit, a mesoscopic point of view, valid, in prin-
ciple, for any of the above descriptions. We consider
the transition between two states (the B-DNA and the
overstretched state) using a stochastic equation of motion
with an asymmetric double-well potential in the presence
of one free parameter connected to biophysical proper-
ties of the DNA chain and/or of the surrounding sol-
vent. The model is able to reproduce the equilibrium
(over)stretching curve with good precision. Moreover,
the approach used here allows us to study the dynam-
ics of the transition. It can be used to carry out real-
istic calculations in DNA dynamics, as in translocation
under strong forces caused by molecular and nanotech-
nological motors, and represents a valid tool to perform
out-of-equilibrium modeling. In this sense the model can
4
be expanded in order to include other variables able to
describe more specific microscopic pictures, such as, for
example, the twist-stretch coupling previously discussed,
or melting properties.
Acknowledgments. This work has been supported by
the Spanish DGICYT Projects No. FIS2008-01240 and
No. FIS2011-25167, co-financed by Fondo Europeo de
Desarrollo Regional (FEDER) funds. Financial support
from European Science Foundation Research Network
"Exploring the Physics of Small Devices" is acknowl-
edged. We also want to thank Professor Alessandro Peliz-
zola for useful conversations.
[1] S. B. Smith, L. Finzi, C. Bustamante, Science 258, 1122
100 018106 (2008).
(1992).
[19] S.Cocco, J. Yan, J-F Leger, D. Chatenay, and J.F.
[2] J. F. Marko and E. D. Siggia, Macromolecules 28, 8759
Marko, Phys. Rev. E 70 011910 (2004).
(1995).
[20] D. Marenduzzo, E. Orlandini, F Seno, and A. Trovato,
[3] P. J. Flory, Statistical Mechanics of Chain Molecules (In-
Phys. Rev. E 81 051926 (2010).
terscience Publishers, New York, 1969).
[21] M. C. Williams, I. Rouzina, and M. J. McCauley, Proc.
[4] T. Odijk, Macromolecules 28, 7016 (1995).
[5] S. B. Smith, Y. Cui, C. Bustamante, Science 271, 795
(1996).
[6] P. Cluzel, A. Lebrun, C. Heller, R. Lavery, J. L. Viovy,
D. Chatenay, and F. Caron, Science, 271 792 (1996).
Natl. Acad. Sci. USA 106 18047 (2009).
[22] J. van Mamerem, P. Gross, G. Farge, P. Hooijman, M.
Modesti, M. Falkenberg, G. J. L. Wuite, and E. J. G.
Peterman, Proc. Natl. Acad. Sci. USA 106 18231 (2009).
[23] P. Bianco, L. Bonigini, L. Melli, M. Dolfi, and V. Lom-
[7] M. J. Fraser and R. D. B. Fraser, Nature (London) 167
bardi, Biophys. J. 101 866 (2011).
761 (1951).
[24] H. Fu, Hu Chen, J. F. Marko, and Jie Yan, Nucl. Ac.
[8] M. Konrad and J. I. Bolonick, J. Am. Chem. Soc. 118
Res. 38 5594 (2010).
10989 (1996).
[25] X. Zhang, H. Chen, H. Fu, P. S. Doyle, and J. Yan,Proc.
[9] T. R. Strick, J.-F. Allemand, D. Bensimon, A. Bensimon,
Natl. Acad. Sci. USA 109 8103 (2012).
and V. Croquette, Science 271 1835 (1996).
[10] R. D. Kamien, T. C. Lubensky, P. Nelson, and C. S.
[26] P. E. J. Rouse, J. Chem. Phys. 21, 1272 (1953).
[27] J. Maji, and S. M. Bhattacharjee, Europhys. Lett. 92
O'Hern, Europhys. Lett. 38 237 (1997).
58004 (2010).
[11] J. D. Moroz, and P. Nelson, Proc. Natl. Acad. Sci. USA
98 14418 (1997).
[12] J. F. L´eger, G. Romano, A. Sarkar, J. Robert, L. Bour-
dieu, D. Chatenay, and J. F. Marko, Phys. Rev. Lett. 83
1066 (1999).
[13] P. Cizeau and J.L. Viovy, Biopolymers 42 383 (1997).
[14] C. Storm, adn P. Nelson Europhys. Lett. 62 760 (2003).
[15] J. F. Marko, Phys. Rev. E 57 2134 (1998).
[16] M. C. Williams, J. R. Wenner, I. Rouzina, and V. A.
Bloomfield, Biophys. J. 80 874 (2001).
[17] J. R. Wenner, M. C. Williams, I. Rouzina, and V. A.
Bloomfield, Biophys. J. 82 3160 (2002).
[18] A. Hanke, M. G. Ochoa, and R. Metzler, Phys. Rev. Lett.
[28] P. M. Chaikin and T. C. Lubensky, Principles of Con-
densed Matter Physics (Cambridge University Press,
Cambridge, 1995).
[29] S. Whitelam, S. Pronk, and P. L. Geissler, Biophys. J.
94 2452 (2008).
[30] H. Fu, H. Chen, C. G. Koh, and C. T. Lim, Eur. Phys.
J. E 29, 45 (2009).
[31] C. Ke, M. Humeniuk, H. S-Gracz, and P. E. Marszalek,
Phys. Rev. Lett. 99 018302 (2007).
[32] C. Danilowicz, C. Limouse, K. Hatch, A. Conover, V. W.
Coljee, N. Kleckner, and M. Prentiss, Proc. Natl. Acad.
Sci. USA 106 13196 (2009).
|
1705.01537 | 1 | 1705 | 2017-05-02T20:45:53 | Absence of selection for quantum coherence in the Fenna-Matthews-Olson complex: a combined evolutionary and excitonic study | [
"physics.bio-ph",
"physics.chem-ph"
] | We present a study on the evolution of the Fenna-Matthews-Olson bacterial photosynthetic pigment-protein complex. This protein complex functions as an antenna. It transports absorbed photons - excitons - to a reaction center where photosynthetic reactions initiate. The efficiency of exciton transport is therefore fundamental for the photosynthetic bacterium's survival. We have reconstructed an ancestor of the complex to establish whether coherence in the exciton transport was selected for or optimized over time. We have also investigated on the role of optimizing free energy variation upon folding in evolution. We studied whether mutations which connect the ancestor to current day species were stabilizing or destabilizing from a thermodynamic view point. From this study, we established that most of these mutations were thermodynamically neutral. Furthermore, we did not see a large change in exciton transport efficiency or coherence and thus our results predict that exciton coherence was not specifically selected for. | physics.bio-ph | physics | Absence of selection for quantum coherence in the Fenna-Matthews-Olson complex: a
combined evolutionary and excitonic study
Stéphanie Valleau,1 Romain A. Studer,2 Florian Häse,1 Christoph Kreisbeck,1 Rafael
G. Saer,3 Robert E. Blankenship,3 Eugene I. Shakhnovich,1 and Alán Aspuru-Guzik1
1Department of Chemistry and Chemical biology,
Harvard University, Cambridge, MA, 02138, USA.∗
2European Bioinformatics Institute (EMBL-EBI),
Wellcome Genome Campus, Hinxton, Cambridge, CB10 1SD, UK.
3Departments of Biology and Chemistry, Washington University in Saint Louis,
One Brookings Drive, St. Louis, Missouri 63130 United States
7
1
0
2
y
a
M
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
3
5
1
0
.
5
0
7
1
:
v
i
X
r
a
We present a study on the evolution of the Fenna-Matthews-Olson bacterial photosynthetic
pigment-protein complex. This protein complex functions as an antenna. It transports absorbed
photons -excitons -to a reaction center where photosynthetic reactions initiate. The efficiency of
exciton transport is therefore fundamental for the photosynthetic bacterium's survival. We have re-
constructed an ancestor of the complex to establish whether coherence in the exciton transport was
selected for or optimized over time. We have also investigated on the role of optimizing free energy
variation upon folding in evolution. We studied whether mutations which connect the ancestor to
current day species were stabilizing or destabilizing from a thermodynamic view point. From this
study, we established that most of these mutations were thermodynamically neutral. Furthermore,
we did not see a large change in exciton transport efficiency or coherence and thus our results predict
that exciton coherence was not specifically selected for.
The Fenna-Matthews-Olson complex is a light-
harvesting protein complex found in green sulfur bacte-
ria. These bacteria first appeared about 1.6 bilion years
ago, in the Proterozoic era [1]. Green sulfur bacteria
survive through anoxygenic photosynthesis; they use sul-
fide and other reduced sulfur compounds or hydrogen as
photosynthetic electron donors [2].
In the last decade, much research has been dedicated
to understanding the excitation energy transfer in the
Fenna-Matthews-Olson (FMO) protein complex [3–13].
Once absorbed, photons become molecular excitations,
or excitons. These excitons are then transported through
the complex due to their interactions with neighboring
excitations, the protein scaffold and the environment. A
renewed interest in this complex arose when low tempera-
ture 2D-spectroscopy experiments suggested the presence
of quantum coherence effects [14, 15] in the exciton dy-
namics. Quantum coherent effects can be thought of as
concurrent beats between electronic excitations which do
not occur classically. Following these experiments, much
theoretical work was carried out [6, 16–19] with the goal
of understanding whether quantum effects were present
and if so, how they contributed to the exciton transport.
Most of the studies have relied on the existence of an
X-ray crystal structure of the FMO complex of either
Chlorobaculum tepidum or Prosthecochloris aestuarii.
FMO is found in all of the anaerobic Chlorobi phyla
and recently it has also been found in aerobic Chloraci-
dobacterium thermophilum of the Acidobacteria phyla
[20].
In these organisms, the FMO complex forms a
homo trimer (see Fig. 1). Each monomer consists of 7/8
∗ [email protected]
bacteriochlorophyll-a (BChl-a) pigment molecules en-
closed in a protein scaffold. The BChl pigment molecules
interact with the protein scaffold through coordination
and hydrogen bonds. The complex is thought to act
as an excitonic wire, funneling the excitation from the
baseplate to the reaction center where charge separation
occurs. This charge separation enables reactions which
are fundamental to the organisms' survival. The biolog-
ical origin of the complex remains a mystery. Olson et
al. (Ref. [21]) speculated that the FMO complex might
have come from an ancient reaction center. They looked
for homology between the FMO protein and PscA, the
reaction center protein of green sulfur bacteria. They
found a signature sequence common to both, however
the sequences only had a 13% identity score. More re-
cently, the role of some specific site mutations [22] and
of cysteines in the FMO protein has been investigated as
well [23]. In Ref. [23] it was found that cysteines appear
to be fundamental for the photoprotection mechanism of
the protein complex.
Regarding exciton transport in this system, there have
only been few efforts to understand whether and how
protein evolution might have influenced it. For instance,
in Ref. [24], the authors computed the BChl transition
energies and looked at the effect of including the point
charges coming from the protein. They also looked at
how changes in the polar groups of amino acids, such
as the amino or hydroxyl groups, influence BChl transi-
tion energies. They found that the charges coming from
the alpha helices seem to influence the energies strongly.
On the other hand, no large change in the energies was
observed when they modified the charge distribution of
single amino acids as a means to emulate single point
mutations. Experimentally, the Blankenship group [25]
looked at comparing the optical properties of FMO com-
2
mutant −∆Gf old
Figure 2. Histogram of free energy variation upon folding,
∆∆G = ∆Gf old
wild, for all possible single-point mu-
tations in the FMO complex trimer of P. aestuarii (PDB:
3EOJ). The coloring is shaded to be more red for destabilizing
mutations (positive values) and blue for stabilizing mutations
(negative values). The bins in the histogram are normalized
by the area under the histogram.
ence in exciton transport and thermodynamic stability
(free energy change upon folding) guided the evolution
of the protein complex. Hence, we carried out calcula-
tions which trace evolution of exciton transport as well
as evolution of thermodynamic stability. For the ther-
modynamics, we studied the stability of the current day
protein complex to mutations by computing and compar-
ing the folding free energy change of the mutant to that of
the non mutated protein complex. Subsequently, we con-
structed an ancestor for the FMO complex and computed
the folding free energies due to point mutations along the
phylogenetic tree. Thus, with these two results, we were
able to compare the stability of C. tepidum to the sta-
bility of it's ancestors and see how it changed along the
tree. To investigate on role of exciton transport, we com-
puted and compared exciton transport in the ancestor
with that of C. tepidum. Because two-dimensional pho-
ton echo (2D-echo) spectroscopy is the experimental tool
employed to reveal coherent phenomena, we computed
this spectrum to determine the role of coherence in the
evolution from ancestor to C. tepidum. Finally, from the
reconstructed phylogenetic tree, we identified sites un-
der positive selection and determined whether positively
selected amino acids were close to BChl molecules and
could influence exciton transport.
RESULTS
Robustness of current day FMO proteins to mutations
We computed the free energy change upon folding,
, as a measure of robustness
∆∆G = ∆Gf old
of the FMO complex to mutations. Calculations were
carried out using FoldX [26] for the trimers of C. tepidum
mutant − ∆Gf old
wild
Figure 1. In panel A, an image describes the location of the
FMO complex in the light-harvesting complex of green sul-
fur bacteria. The photons are absorbed by the chlorosome
and transported as excitations to the FMO complex and ul-
timately to the reaction center. In panel B one can see crys-
tal structure of FMO complex monomer for Chlorobaculum
tepidum (PDB: 3ENI). The protein scaffold is colored in pink
for alpha helices, in yellow for beta sheets, and in blue for
loops. The 8 bacteriochlorophyl-a (BChl-a) pigments are col-
ored in cyan (their side chains are not shown for simplicity).
ing from three different species to understand the effect of
the protein scaffold. They observed spectral differences,
and these were assigned to the way the BChl molecules
bind to the protein scaffold in each species. This suggests
that one may probe exciton dynamics and the biologi-
cal role of coherence through point mutations.
In this
sense, we are most interested in mutations occurring in
the vicinity of BChl binding sites as the effect of their
charge is expected to have a stronger influence on the
BChl energies.
For the above reasons, we have carried out a theoretical
study of the protein complex from an evolutionary per-
spective. We aimed to identify whether quantum coher-
Reaction centerBaseplateChlorosomeFMOFMOCytoplasmPeriplasmFMO
monomerLight-harvesting
complexABdestabilizingstabilizingProportionand P. aestuarii for all possible single point mutations
(see Methods for details). In our model, each single point
mutation occurs at the same time in all monomers of the
trimer. We found that most single point mutations would
be destabilizing ~ 60% (see Fig. 2). The overall landscape
of mutations is the same for C. tepidum and P. aestuarii
(see Fig. S2 in the Supporting Information). This is qual-
itatively similar to the general statistics of the effects of
mutations on stability (see Ref. [27, 28]).
We also looked at how the free energy change varies
depending on the location of the mutation in the pro-
tein complex monomers of P. aestuarii and C. tepidum
and found that mutations at the interface of monomers
are among the most destabilizing (see Fig. S3).These ob-
servations rely on the assumption that each single point
mutation (in each monomer) does not impede the assem-
bling of a trimeric structure.
Ancestor reconstruction, selection of mutations and
variation of the free energy landscape
A phylogenetic tree was generated using the multiple
sequence alignment built from the fmoA amino acid se-
quences (see details in Methods). The amino acid se-
quences were obtained by translating the fmoA monomer
gene sequences. In Fig. 3 we show the cladogram tree
(phylogenetic tree is also shown in Fig. S1). The tree
divides in two main clades: saltwater (e.g. P. aestuarii)
and freshwater (e.g. C. tepidum) bacteria.
At each node of the tree, the sequence and structure
of the protein scaffold were reconstructed. The ances-
tral amino acid sequences were obtained using Bayesian
inference and the scaffold structures were computed us-
ing homology modeling (see Methods). The pie-charts
above each branch indicate, based on the diameter, how
many mutations occurred between the two branch nodes,
and based on the color, whether they were stabilizing
(blue), neutral (grey) or destabilizing (red). Most muta-
tions were neutral in terms of modifying the free energy
change upon folding.
Given the inferred ancestor amino acid sequence at
the root we constructed a guess for the full ancestral
structure (protein scaffold and BChl molecules) through
a combination of homology modeling and molecular dy-
namics simulations (see Methods). An image of the struc-
ture of the putative ancestor compared to that of C.
tepidum for one monomer is shown in Fig. 4. The im-
ages in each panel are constructed from a snapshot of
the molecular dynamics simulations at 300K after equi-
libration. In panel B, we show two orientations of one
monomer of the protein complex. No large difference in
the backbone of the protein is observed. This is due to
the fact that the protein scaffold was constructed with
homology modelling. The differences come from the lo-
cation of side chains and the change in charge of the mu-
tated amino acids. In panel B, mutations are highlighted
using a color scheme based on the BLOSUM90 distance.
3
We see a few red mutations (large negative BLOSUM90
distance) on the beta sheet which could be a problem
for the formation of the trimer. Another effect of the
mutations is the change in orientation of BChl molecules
(see Panel A). This change in orientation of the BChls
influences the coupling amongst their first excited states
(see Table S1). In particular, the coupling between the
first excited states of BChl 3 and 4 is weaker in the an-
cestor than in C. tepidum and, on the other hand, the
coupling between the first excited states of BChl 4 and
5 is stronger in the ancestor than it is in C. tepidum.
Finally, as can be seen in Tab. S2, the average distance
of the amino acids from the BChl molecules changes be-
tween the ancestor and C. tepidum and the differences
are on the order of 0.1-0.4Å.
Given the phylogenetic tree, we also identified branches
under positive selection (bright green) by using the
branch-site model (details in Methods).
In Fig. 5, we
show mutations which were positively selected for along
these branches in the structure of C. tepidum. Most
positively selected mutations are distant from the BChl
residues, and therefore we currently do not believe they
had a strong role in modifying exciton transport. Given
the high computational cost associated with constructing
the full structure (with BChl molecules) at each node,
we did not investigate the role of each positively selected
mutation using atomistic ab initio methods. However,
we computed the free-energy changes associated with the
mutations.
In Fig. 6, panel A, we show the histogram of free en-
ergy changes which connects the ancestor to current day
species in red and, in cyan, the histogram of folding free
energy variation for all possible single-point mutations in
the FMO complex trimer of C. tepidum. We found that
65% of mutations are destabilizing for C. tepidum. On
the other hand only 32% of mutations of the ancestral
FMO complex along the phylogenetic tree were desta-
bilizing while 21% were stabilizing and 47% were neu-
tral. This could indicate that mutations which favored
the folding free energy change variation during evolution
led to more stable current day protein structures. How-
ever, we cannot claim this in general as the trend changes
for different current day species. For instance, as can be
seen by comparing the two figures in panel B of Fig. 6,
the distribution of folding free energy changes for muta-
tions which connect the ancestor to P. aestuarii is signif-
icantly different from that for mutations which connect
the ancestor to C. tepidum. This difference was con-
firmed by the Kolmogorov-Smirnov test. Furthermore,
we computed that, for the mutations which connect the
ancestor to P. aestuarii, 25 % are stabilizing and 19% are
destabilizing. However, in the case of mutations which
lead to C. tepidum, 47% are destabilizing while 22% are
stabilizing. This indicates that C. tepidum is less stable
than the ancestor while P. aestuarii is more stable than
the ancestor.
4
Figure 3. Rooted cladogram for fmoA computed using MAFFT and TranslatorX [29] for the alignment and PHYML [30].
Above each branch we show a pie-chart label with a diameter proportional to the number of single point mutations and color
based on the associated free energy change ∆∆G. Red represents destabilizing mutations, blue represents stabilizing mutations,
and grey represents neutral mutations.
Exciton transport and coherence
Absorption, circular dichroism and linear dichroism
In Fig.
7 we compare the absorption spectrum of
the ancestor to that of C. tepidum. The ancestor spec-
trum shows three main absorption peaks at 807, 819 and
830 nm while C. tepidum shows strong absorption at 804,
814 and 826 nm. These trends and the corresponding cir-
cular dichroism spectra (see Supporting Information) of
the ancestor and of C. tepidum are very similar, respec-
tively, to the spectra of FMO in C. thiosulfatophilum and
C. limicola reported in Ref. [31]. However, we do not see
strong evidence of a similarity between the ancestor spec-
trum and the reaction center absorption spectra reported
in that work. The small bump at about 845 nm (see Fig.
7) is most likely due to noise, as it does not correspond
to any of the frequencies of the equivalent Fermi-Golden
rule absorption spectrum. Therefore, our current results
do not confirm that the ancestor is related to current day
PscA, the reaction center protein of green sulfur bacteria,
however we cannot exclude that it may have been related
to ancestral reaction centers.
In the Supporting Information we also show the com-
parison of our C. tepidum spectra to the experimental
absorption, linear dichroism and circular dichroism spec-
5
Figure 4. In panel A, we show the structure of the BChl molecules in a monomer of the ancestral structure (yellow) overlaid
with the BChl molecules in the current day structure of C. tepidum (light grey). The images were generated from snapshots of
the molecular dynamics simulations at 300K. In panel B, for the same snapshots we compare the backbone of the ancestral FMO
structure (yellow) to that of the current day C. tepidum structure (light grey backbone). Some amino acids are colored based
on the BLOSUM90 matrix, which quantifies how different they are (positive numbers indicate easy substitution while negative
numbers indicate difficult substitution). These amino acids are also highlighted in the legend in panel C. The structures are
snapshots from the molecular dynamics simulations at 300K. The amino acids which differ in the two structures are colored
based on the BLOSUM90 matrix which quantifies how likely it is to substitute one amino acid with the other. The amino acid
mutants in the ancestral structure are also shown using the bond and atom representation.
tra.
Coherence and 2D spectra
Using the hierarchy equation of motion (HEOM) non-
Markovian master equation method [32–34], we com-
puted the 2D-echo spectra of the reconstructed ances-
tral FMO complex and of C. tepidum at 150 K. To
this end we employed the QMaster software package [35]
(for more details see Methods). The spectral densities
which were employed are described in Methods. These
were obtained by fitting the original QM/MM spectral
densities to Drude-Lorentz spectral densities with three
peaks. The spectral densities indicated a similar cou-
pling strength of the system to the bath for the ancestral
and C. tepidum FMO complexes for almost all modes.
The total signal of the 2D-echo spectra, i.e. sum of the
stimulated emission (SE), ground state bleaching (GB),
and excited state absorption (ESA) pathways is shown
in Fig. 8. From the oscillations of the cross-peaks in
panel B) we cannot deduce an improvement in quantum
coherence or a significant difference between the ancestor
and C. tepidum. For all species, cross-peak beatings in
the 2D-echo spectra are dominated by ground state vi-
Figure 5. Positively selected sites shown in the structure of C.
tepidum (PDB: 3ENI). The color gradient corresponds to the
strength of positive selection (darkest red - strongest selection,
blue - no positive selection). We only plot the side chains
of positively selected residues which are within 6 Å of the
magnesium in the BChl molecules.
6
brations [36] superposed by minor vibronic contributions.
This suggests that optimizing exciton transport through
quantum coherence was not part of the evolution of the
FMO complex.
CONCLUSION
We have reconstructed a structure for the ancestral
Fenna-Matthews-Olson complex and found that muta-
tions of it's amino acids over time have influenced the free
energy upon folding of the protein complex. In particu-
lar, the current day specie of P. aestuarii was obtained
through stabilizing mutations while that of C. tepidum
is less stable than the ancestor. Regarding exciton trans-
port, we do not observe a significant change or improve-
ment in the efficiency of exciton transport or quantum co-
herent transport. These results suggest that the complex
did not evolve to optimize for quantum coherent trans-
port.
It is also possible that the ancestor was already
sufficiently optimized for exciton transport and therefore
further evolution or optimization was not necessary. One
of the results we find is that mutations most likely led to
better binding of the FMO complex to the baseplate and
reaction center. This is confirmed by the larger instabil-
ity associated with mutating residues in those locations.
This leaves an open question: perhaps the overall exci-
ton transport, from the chlorosome to the reaction center,
was optimized by favoring mutations in FMO which led
to a stronger binding of FMO to its neighbouring protein
systems.
For future work, we are interested in looking at the
exciton transport at each node of the phylogenetic tree
and further in expressing the ancestral sequence in cur-
rent day C. tepidum to obtain an experimental structure
and 2D-echo spectra for the ancestor. Finally it would
be interesting to look at the role of photoprotection to
see whether the mutations could have influenced it.
METHODS AND COMPUTATIONAL
DETAILS
Stability of the C. tepidum and P. aestuarii
Fenna-Matthews-Olson complex to mutations
Each of the amino acids of the trimer and monomer of
the FMO complex of P. aestuarii and C. tepidum were
mutated to the each of the other possible 19 amino acids
using FoldX [26]. The corresponding free energy varia-
tion upon folding ∆∆G = ∆Gmut − ∆Gwt was obtained
for each mutation. In order to account for the effect of
mutations at the interface between monomers within the
trimer, the mutations were carried out simultaneously in
each monomer of the trimer and the resulting free energy
variation was normalized by dividing by three.
In FoldX, mutations are modeled as follows. The ini-
tial crystal structure is optimized to remove any eventual
Figure 6. Panel A: In blue we show a histogram of the folding
free energy changes for all possible single point mutations of
the current day C. tepidum FMO complex structure. In red,
we show a histogram of the folding free energy changes for
all the single point mutations which are necessary to connect
the ancestral FMO structure to the current day structures
(i.e. all single point mutations along all branches). Panel B:
Right-hand side: histogram of the folding free energy changes
for the mutations which connect P. aestuarii to the ancestor.
Left-hand side: histogram of the folding free energy changes
for the mutations which connect C. tepidum to the ancestor.
Note that, for all histograms, each bin contains the number
of points with energy in that interval normalized by the area
under the entire histogram.
Figure 7. Comparison of simulated absorption spectra of the
ancestral FMO complex (blue line) to that of current day
C. tepidum at 300K (grey dashed line) in arbitrary units.
Spectra were computed using the Qy transitions obtained by
using QM/MM. See Methods for more details.
7
Figure 8. Panel A: Contour plots of the total computed signal for the 2D-echo spectra at different delay times. The figures
in the top row show results for the ancestor at a delay time of 90 and 400 fs and the bottom figures show the analogous plots
for C. tepidum. Panel B: The amplitude of oscillation of two specific cross peaks. The top image is for the ancestor and the
bottom image is for C. tepidum.
steric clashes. Then, the residue of interest is mutated
and its nearest neighbors are mutated to themselves and
conformationally relaxed to remove any local clash. The
nearest neighbors are mutated to themselves so that their
geometry may be optimized toghether with that of the
central residue. During this procedure the backbone of
the protein is kept fixed and all other residues that are far
from the one of interest are also kept fixed. The stabil-
ity, ∆Gwt, of this relaxed structure is obtained by using
an effective energy function. Subsequently, the residue of
interest is mutated to each of the other 19 amino acids
and its neighbors to themselves and the various ∆Gmut
are computed.
The procedure described above was repeated 5 times
for each mutation to ensure that the minimum energy
conformations of large residues that have many rotamers
were identified. The effective energy function ("effective
energy" here refers to the Helmoltz free energy of a sys-
tem (protein + solvent) for a fixed protein conformation)
in FoldX has been optimized for amino acid sequences,
thus all BChl-a molecules could not be included directly
in this calculation. To account for their presence, the
structures given in input to FoldX were obtained by ho-
mology modeling with Modeller [37].
In the Modeller
simulations, the BChl molecules were inserted and kept
fixed as hard spheres.
In the FoldX optimization the
residues that are known to bind to BChl-a molecules
from the crystal structure analysis of FMO complexes
from current day species were also kept fixed. A similar
procedure was recently employed successfully for the case
of RubisCO where Mg atoms are present [38].
Determination of positively selected sites: site and
branch-site models
Phylogenetic tree
All fmoA gene sequences from the EMBL database
were gathered and employed to generate an alignment
using translatorX with MAFFT [29]. Phylogenetic trees
were constructed using PhyML [30] with 5000 bootstraps.
The parameters and settings used in PhyML were as fol-
lowing. The LG substitution model was chosen. We
selected to have four substitution rate categories. The
alpha parameter for the gamma distribution of sites was
set to be estimated by the code. Both NNIs and SPR
methods were used to search for the optimal tree topol-
ogy and finally, tree topology, branch lengths and rate
parameters were chosen to be optimized by the code. A
set of three trees were constructed. In the first tree, all
sequences were included excluding FJ210646, as it was
missing about 100 residues. Therefore, including it would
have introduced more error in the tree reconstruction
phase. For the second tree, we removed Chloroacidobac-
terium thermophilum (ABV27353) as it had the lowest
sequence identity percent and thus also introduced more
uncertainty. In the third tree, Chloroherpeton thalassium
(ACF13179) was removed as it was the second most dis-
tant sequence from the rest. We computed ancestral se-
quences and searched for positively selected sites on all
three trees but we present results for the third tree as we
believe it has the smallest error.
Site and branch-site models
The branches under positive selection were identi-
fied using the branch-site model [39] as implemented in
CodeML. This model allows for variation of ω = dN
dS
(the ratio of synonymous to non synonymous mutations)
amongst branches and sites. For each simulation, one
branch is selected as a foreground branch and, using
CodeML, we computed the likelihood that the branch
was under positive selection. The input DNA sequence
alignment was the same as the one generated to deduce
the phylogenetic tree. Given a branch under positive
selection, we determined which amino acids were under
positive selection by using the Bayes Empirical Bayes
(BEB) model [40]. We corrected the probability values
p-values for False Discovery Rate (FDR) by using the
q-value [41], as discussed in Ref. [42].
Reconstruction of the ancestral protein structure of FMO
Sequence reconstruction
An ancestral sequence was reconstructed using max-
ium likelihood as implemented in FastML [43] and the
LG+G model for amino acid substitution. The LG+G
model was established to be the best model under the
Bayesian Information Criterion (BIC). With FastML, we
obtained the most probable sequences, together with the
posterior probabilities for each character and indel at
each sequence position for each internal node of the tree.
As inputs, we used the phylogenetic trees described in
the previous section and rooted them based on the most
distant sequence. The alignments used in the input were
the same as those employed to generate the trees. For the
last tree that excludes C. thermophilum and C. thalas-
sium, we used the midpoint root, as most sequences are
similar, (see Fig. 3) and the branching is almost identical
to that obtained when C. thalassium or C. thermophilum
are included.
8
quence using its alignment to the related protein struc-
tures as a guide. The restraints are obtained based
on the assumption that the structures will be similar
in the aligned regions.
In MODELLER, the form of
the restraints was obtained from a statistical analysis
of the relationships between similar protein structures
in a database of 105 alignments that included 416 pro-
teins of known 3D structure [37]. These restraints are
supplemented by stereochemical restraints such as bond
lengths, angles etc, which are obtained from a force field.
Once all these restraints are established, the model struc-
ture is obtained in Cartesian space by minimizing the
violations of all restraints.
We used the structure of P. aestuarii (PDB: 3EOJ)
as the homologue and generated 100 possible structures
with slow refinement. The ancestral structure was chosen
to be the one with the best molpdb factor. The initally
reconstructed structure only included the protein scaf-
fold, as there is no current parametrization for BChl's
in Modeller. Thus, BChl molecules were included sub-
sequently using minimization as implemented in NAMD
[44]. The chromophores were initially positioned as in the
homologous structure and the structure was optimized by
minimizing the energy in two steps. In the first step, the
backbone was kept fixed, and for the second minimiza-
tion the backbone was free to move. The structure was
subsequently equilibrated without constraints using the
Amber14 software package, [45] with the Amber ff99SB
force field[46]. The BChl-a parameters employed are re-
ported in Refs. [11, 47]. The protonation states of all
amino acids were determined with the H++ 3.0 software
[48], under neutral pH conditions. All complexes were
solvated using TIP3P periodic water boxes,
[49] with
a minimum distance of 15 Å between the complex and
the box boundaries. Charges were neutralized by adding
sodium ions for the ancestor and P. aestuarii and chlo-
ride for C. tepidum. Shake constraints were applied to
all bonds containing hydrogen. Minimizations were car-
ried out for 2000 steps for P. aestuarii and C. tepidum
and for 10000 steps for the ancestor. Minimizations were
followed by 200 ps adaptation runs to impose a tempera-
ture of 300 K and a pressure of 1 atm on the systems. All
three complexes were equilibrated for 50 ns in the same
environmental conditions. Long range electrostatic in-
teractions were calculated using the Particle-Mesh Ewald
method [50]. For each complex we then carried out 40 ps
production runs with a 1 fs integration step.
Structure reconstruction and molecular dynamics
Correlating biological stability and evolution to exciton
dynamics
The ancestral structure was built by using homology
modeling with satisfaction of spatial restraints as imple-
mented in Modeller [37]. The method of modeling by
satisfaction of spatial restraints works as follows. Con-
straints are generated on the structure of the target se-
QM/MM Hamiltonian
For the exciton dynamics, the FMO complex is simu-
lated as a system coupled to a bath (the protein environ-
9
order response function S(3)(t1, Tdelay, t3). Phase match-
ing ensures that the components of the detected signal go
along distinct directions, and thus can be experimentally
separated. We considered the rephasing (RP) component
of the signal
(2)
IRP (ω1, Tdelay, ω3) = ...
∞
dt1dt3eiω3t3−iω1t1 S
(3)
RP (t1, Tdelayt3)
0
for which coherent phenomena show as unique oscillatory
pattern in the cross-peak dynamics as a function of de-
lay time. Using the notation of double sided Feynman
diagrams [55, 56], the total signal is given by three sep-
arate pathways: stimulated emission (SE), ground state
bleaching (GB) and excited state absorption (ESA). The
latter involves the double exciton manifold. We numer-
ically evaluated IRP (ω1, Tdelay, ω3) by propagating the
reduced density matrix, describing the exciton degrees
of freedom, along the double sided Feynman diagrams.
Hereby, each interaction with the laser pulses requires an
interruption of the dynamics and a multiplication of the
reduced density matrix with the dipole operator, either
from the right or from the left. The propagation was
done using QMaster [35], a high-performance implemen-
tation of HEOM [57, 58]. The rotational average over
random orientations of the probed sample was included
by sampling 20 laser polarization vectors aligned with the
vertices of a dodecahedron [54].
ACKNOWLEDGEMENTS
S. V. would like to thank Pouria Dasmeh and Adrian
Serohijos for their interesting discussions on methods to
compute the free energy changes for mutations and phy-
logenetic tree reconstruction. The authors thank Nicolas
Sawaya for sharing his code for the calculation of atomic
partial charges from the transition densities. The com-
putations in this paper were completed on the Odyssey
cluster supported by the FAS Division of Science, Re-
search Computing Group at Harvard University. S. V.,
F. H. and A. A.-G. acknowledges support from the Cen-
ter for Excitonics, an Energy Frontier Research Center
funded by the U.S. Department of Energy, Office of Sci-
ence and Office of Basic Energy Sciences under Award
Number de-sc0001088. A.A.G. acknowledges the gener-
ous support from the Canadian Institute for Advanced
Research. This work was supported as part of the Photo-
synthetic Antenna Research Center (PARC), an Energy
Frontier Research Center funded by the U.S. Department
of Energy, Office of Science, Office of Basic Energy Sci-
ences under Award Number de-sc 0001035 to REB. RS
was supported by the PARC grant.
Figure 9. Average first excited state energies hEnifor each
BChl molecule in the FMO complex of C. tepidum (black),
P. aestuarii (grey) and for the ancestral structure (blue). Val-
ues were obtained by averaging over the production QM/MM
trajectory at 300K.
ment). The system Hamiltonian is defined as
8X
i=1
Hex =
8X
i6=j
εiiihi +
Vijiihj
(1)
where εi is the first excited state energy of the i-th BChl.
The Vij terms correspond to the excitonic coupling be-
tween the excited states for the i−th and j−th molecule.
Site energies were computed using TDDFT in QChem
[51] with the PBE0 functional [52] and 3-21G basis set
for 10000 frames (40 ps) of the MD production run. Cou-
plings were calculated using two approaches: the point-
dipole approximation (PDA) and via transition charges
from the electrostatic potential (TrEsp) [53]. TrEsp has
shown to be the most accurate, although it is computa-
tionally more demanding than using the PDA [53]. The
computed couplings are given in the Supplementary in-
formation.
The average first excited state energy of each BChl
isshown in Fig. 9. We notice there is no large change in
the trend of the excited state energies for the BChl's.
Exciton coherence and 2d spectrum
[54].
The 2d-echo spectra in Fig. 8 were calculated as in
Ref.
In two-dimensional electronic spectroscopy
the sample is probed by a sequence of thee laser pulses.
Adjusting the delay time Tdelay between the second and
the third pulse yields a time resolved picture of the ex-
citon dynamics. Within the impulsive limit (assuming
δ-pulses), the 2D-echo spectrum is related to the third
[1] J. J. Brocks, G. D. Love, R. E. Summons, A. H. Knoll,
G. A. Logan, and S. A. Bowden, Nature 437, 866 (2005).
[2] R. E. Blankenship,
2014).
(Wiley-Blackwell, Oxford, UK,
[3] J. Adolphs and T. Renger, Biophys. J. 91, 2778 (2006).
[4] M. Mohseni, P. Rebentrost, S. Lloyd, and A. Aspuru-
Guzik, J. Chem. Phys. 129, 174106 (2008).
[5] P. Rebentrost, M. Mohseni, I. Kassal, S. Lloyd,
and
[6] A. Ishizaki and G. R. Fleming, Proc. Natl Acad. Sci.
A. Aspuru-Guzik,.
USA. 106, 17255 (2009).
[30] S. Guindon, J.-F. Dufayard, V. Lefort, M. Anisimova,
W. Hordijk, and O. Gascuel, Systematic Biology 59,
307 (2010).
[31] J. Olson, K. Bacon, and K. H. Thompson,.
[32] Y. Tanimura and R. Kubo,.
[33] A. Ishizaki and Y. Tanimura,.
[34] A. Ishizaki and G. R. Fleming, J. Chem. Phys. 130,
[7] P. Rebentrost, M. Mohseni, and A. Aspuru-Guzik, J.
234111 (2009).
10
Phys. Chem. B. 113, 9942 (2009).
[8] G. Panitchayangkoon, D. Hayes, K. A. Fransted, J. R.
Caram, E. Harel, J. Wen, R. E. Blankenship, and G. S.
Engel, Proc. Natl Acad. Sci. USA. 107, 12766 (2010).
[9] P. Rebentrost, S. Shim, J. Yuen-Zhou, and A. Aspuru-
Guzik, Procedia Chemistry 3, 332 (2011).
[10] G. Ritschel, J. Roden, W. T. Strunz, A. Aspuru-Guzik,
and A. Eisfeld, J. Phys. Chem. Lett. 2, 2912 (2011).
Guzik, Biophys. J. 102, 649 (2012).
[12] S. Valleau, A. Eisfeld, and A. Aspuru-Guzik, J. Chem.
[13] J. Dostlál, J. Pśenćík, and D. Zigmantas, Nat. Chem.
(2007).
Phys. 137, 224103 (2012).
(2016), 10.1038/nchem.2525.
[11] S. Shim, P. Rebentrost, S. Valleau,
and A. Aspuru-
[40] Z. Yang, W. S. Wong, and R. Nielsen, Mol. Biol. Evol.
[35] C. Kreisbeck, T. Kramer, and A. Aspuru-Guzik,.
[36] C. Kreisbeck, T. Kramer, and A. Aspuru-Guzik, J. Phys.
Chem. B 117, 9380 (2013).
[37] A. Sali and T. L. Blundell,.
[38] R. A. Studer, P. A. Christin, M. A. Williams, and C. A.
Orengo, Proc. Natl. Acad. Sci. 111, 2223 (2013).
[39] J. Zhang, R. Nielsen, and Z. Yang, Mol. Biol. Evol. 22,
2472 (2005).
22, 1107 (2005).
[41] D. A. Bass JDSwcfAJ and R. D,.
[42] M. Anisimova and Z. Yang, Mol. Biol. Evol. 24, 1219
[43] H. Ashkenazy, O. Penn, A. Doron-Faigenboim, O. Cohen,
G. Cannarozzi, O. Zomer, and T. Pupko, Nucleic Acids
Res. 40, W580 (2012).
[44] J. C. Phillips, R. Braun, W. Wang, J. Gumbart,
E. Tajkhorshid, E. Villa, C. Chipot, R. D. Skeel,
L. KalÃľ, and K. Schulten, Journal of Computational
Chemistry 26, 1781 (2005).
[45] D. A. Case, V. Babin, J. T. Berryman, R. M. Betz,
Q. Cai, D. S. Cerutti, T. E. Cheatham III, T. A. Dar-
den, R. E. Duke, H. Gohlke, A. W. Goetz, S. Gusarov,
N. Homeyer, P. Janowski, J. Kaus, I. Kolossváry, A. Ko-
valenko, T. S. Lee, S. LeGrand, T. Luchko, R. Luo,
B. Madej, K. M. Merz, F. Paesani, D. R. Roe, A. Roit-
berg, C. Sagui, R. Salomon-Ferrer, G. Seabra, C. L. Sim-
merling, W. Smith, J. Swails, R. C. Walker, J. Wang,
R. M. Wolf, X. Wu, and P. A. Kollman, (2014).
[46] T. E. Cheatham, R. Cieplak, and R. A. Kollman,.
[47] S. Chandrasekaran, M. Aghtar, S. Valleau, A. Aspuru-
Guzik, and U. KleinekathÃűfer, J. Phys. Chem. B 119,
9995 (2015), pMID: 26156758.
[48] R. Anandakrishnan, B. Aguilar, and A. V. Onufriev,.
[49] W. L. Jorgensen and J. D. Madura,.
[50] U. Essmann, L. Perera, M. L. Berkowitz, T. Darden,
H. Lee, and L. Pederson,.
[51] Y. Shao, L. F. Molnar, Y. Jung, J. Kussmann, C. Ochsen-
feld, S. T. Brown, A. T. Gilbert, L. V. Slipchenko,
S. V. Levchenko, D. P. O'Neill, R. A. DiStasio Jr, R. C.
Lochan, T. Wang, G. J. Beran, N. A. Besley, J. M.
Herbert, C. Yeh Lin, T. Van Voorhis, S. Hung Chien,
A. Sodt, R. P. Steele, V. A. Rassolov, P. E. Maslen,
P. P. Korambath, R. D. Adamson, B. Austin, J. Baker,
E. F. C. Byrd, H. Dachsel, R. J. Doerksen, A. Dreuw,
B. D. Dunietz, A. D. Dutoi, T. R. Furlani, S. R. Gwalt-
ney, A. Heyden, S. Hirata, C.-P. Hsu, G. Kedziora,
R. Z. Khalliulin, P. Klunzinger, A. M. Lee, M. S. Lee,
W. Liang, I. Lotan, N. Nair, B. Peters, E. I. Proynov,
P. A. Pieniazek, Y. Min Rhee, J. Ritchie, E. Rosta,
C. David Sherrill, A. C. Simmonett, J. E. Subotnik,
H. Lee Woodcock III, W. Zhang, A. T. Bell, A. K.
Chakraborty, D. M. Chipman, F. J. Keil, A. Warshel,
W. J. Hehre, H. F. Schaefer III, J. Kong, A. I. Krylov,
P. M. W. Gill, and M. Head-Gordon, Phys. Chem. Chem.
Phys. 8, 3172 (2006).
[14] G. S. Engel, T. R. Calhoun, E. L. Read, T.-K. Ahn,
T. Mancal, Y.-C. Cheng, R. E. Blankenship, and G. R.
Fleming, Nature. 446, 782 (2007).
[15] G. D. Scholes, G. R. Fleming, L. X. Chen, A. Aspuru-
Guzik, A. Buchleitner, D. F. Coker, G. S. Engel, R. van
Grondelle, A. Ishizaki, D. M. Jonas, J. S. Lundeen, J. K.
McCusker, S. Mukamel, J. P. Ogilvie, A. Olaya-Castro,
M. A. Ratner, F. C. Spano, K. B. Whaley, and X. Zhu,.
[16] F. Fassioli and A. Olaya-Castro,.
[17] C. Kreisbeck and T. Kramer, J. Phys. Chem. Lett. 3,
2828 (2012).
[18] A. Olaya-Castro, C. Lee, F. Olsen,
and N. Johnson,
Phys. Rev. B. 78, 085115 (2008).
[19] C. Olbrich, T. L. C. Jansen, J. Liebers, M. Aghtar,
J. Strümpfer, K. Schulten,
and
U. Kleinekathöfer, J. Phys. Chem. B 115, 8609 (2011).
[20] J. Wen, Y. Tsukatani, W. Cui, H. Zhang, M. L. Gross,
D. A. Bryant, and R. E. Blankenship, BBA - Bioener-
getics 1807, 157 (2011).
J. Knoester,
[21] J. M. Olson,.
[22] R. G. Saer, V. Stadnytskyi, N. C. Magdaong, C. Good-
son, S. Savikhin,
and R. E. Blankenship, Biochimica
et Biophysica Acta (BBA) - Bioenergetics 1858, 288
(2017).
[23] R. Saer, G. S. Orf, X. Lu, H. Zhang, M. J. Cuneo, D. A.
and R. E. Blankenship, BBA - Bioenergetics
Myles,
1857, 1455 (2016).
[24] F. Müh, M. E.-A. Madjet, J. Adolphs, A. Abdurahman,
B. Rabenstein, H. Ishikita, E.-W. Knapp, and T. Renger,
Proc. Natl. Acad. Sci. U. S. A. 104, 16862 (2007).
[25] D. E. Tronrud, J. Wen, L. Gay, and R. E. Blankenship,
Photosynth. Res. 100, 79 (2009).
[26] R. Guerois, J. E. Nielsen, and L. Serrano, J. Mol. Bio.
320, 369 (2002).
[27] K. B. Zeldovich, P. Chen, and E. I. Shakhnovich, Pro-
ceedings of the National Academy of Sciences 104, 16152
(2007).
[28] N. Tokuriki, F. Stricher, J. Schymkowitz, L. Serrano, and
D. S. Tawfik, Journal of Molecular Biology 369, 1318
(2007).
[29] F. Abascal, R. Zardoya, and M. J. Telford,.
[52] J. P. Perdew, M. Ernzerhof, and K. Burke, J. Chem.
Phys. 105, 9982 (1996).
11
[53] M. E. Madjet, A. Abdurahman, and T. Renger,.
[54] B. Hein, C. Kreisbeck, T. Kramer, and M. Rodriguez,.
[55] S. Mukamel, (Oxford: Oxford University Press, 1999).
[56] M. Cho, (Boca Raton, FL: CRC Press, 2009).
[57] C. Kreisbeck, T. Kramer, M. Rodríguez, and B. Hein,.
[58] C.
(2013),
Kreisbeck
Kramer,
and
10.4231/D3RB6W248.
T.
[59] T. Renger, M. E.-A. Madjet, M. Schmidt am Busch,
J. Adolphs, and F. Müh, Photosynth. Rev. 116, 367
(2013).
[60] B. H. Besler, K. M. Merz, and P. A. Kollman, J. Comput.
Chem. 11, 431 (1990).
[61] E. Sigfridsson and U. Ryde, J. Comput. Chem. 19, 377
(1998).
[62] M. M. Francl, C. Carey, L. E. Chirlian, and D. M. Gange,
J. Comput. Chem. 17, 367 (1996).
[63] S. I. E. Vulto, S. Neerken, R. J. W. Louwe, M. A. de Baat,
J. Amesz, and T. J. Aartsma, J. Phys. Chem. B. 102,
10630 (1998).
SUPPORTING INFORMATION (SI)
Phylogenetic tree
In Fig. S1 we show the phylogenetic tree of the FMO
complex, based on the fmoA gene, with branch lengths
defined in terms of sequence identity.
Thermodynamic results
In Fig. S2 we compare the free energy variation his-
tograms for P. aestuarii and C. tepidum. We notice that
the distributions are not significantly different. This was
also checked by a Kolmogorov-Smirnov goodness-of-fit
test.
Free energy changes due to mutations vary depending
on the location in the protein complex monomers of P.
aestuarii and C. tepidum. We found that mutations at
the interface of monomers are destabilizing. We identify
regions where mutations are more stabilizing (green box)
and regions where they are mostly destabilizing (black
box).
In the top left panel of Fig. S3 we show that
the outer amino acids neighboring the 8-th BChl can be
mutated and this causes no stress / very little stress to
the structure. On the other hand, in the top right panel,
we show that it is much more destabilizing to mutate the
inner alpha-helix (black box), which lies at the interface
with another monomer of the trimer. This could indicate
that, in this region, there was a larger priority to stabilize
the trimeric shape rather than the interaction with the
eight BChl.
Hamiltonian
Site couplings were computed using the point-dipole
approximation (PDA) and the transfer charge from elec-
trostatic potential (TrEsp) method [59].
In the point-dipole approximation, the exciton en-
ergy transfer couplings are computed according to Eq. 3,
where ~dA10 denotes the transition dipole of the 0 → 1 tran-
sition for molecule A. Here R is the distance between the
centers of the molecules A and B. To compute the exci-
tation energy transfer couplings in the PDA we assumed
an effective dipole strength of 37.1 D2, an orientation of
transition dipoles along the axis of the ND and NB ni-
trogen atoms in the BChls and set f = 1.
~dA10 · ~dB01
R3 − 3( ~dA10 · R)(~dB01 · R)
R5
!
(3)
10,01 = f
V PDA
4π0
Excitation energy transfer couplings can also be
computed from atomic transition charges qA
i (1, 0) and
j (0, 1) that fit the electrostatic potentials of transition
qB
charge densities. We calculated atomic transition charges
for all atoms in each of the BChls by fitting the transition
12
densities. The transition density cube files were obtained
using TD-DFT with the 6-31G* basis set and PBE0 func-
tional in QChem [51]. For the fitting, we employed a
method inspired by the Merz-Kollman method [60] and
the CHELP-BOW method [61]. A random set of coor-
dinates were chosen, lying more than 2 Å and less than
8 Å from any atom. All points were weighted equally.
We avoided all points lying within the van der Waals
radius as it has been shown that prioritizing points out-
side of this radius led to charges that better reproduced
the electron density's higher multipoles [61]. A sampling
frequency of 6000 points per atom was used.
We used the pseudoinverse method to solve the final
system of equations and singular value decomposition to
find the pseudo inverse. This approach avoids having to
deal with poorly conditioned intermediate matrices [62].
Excitation energy transfer couplings was then com-
puted by summing over the Coulomb potentials between
all pairs of atomic transition charges in the two molecules
(see Eq. 4)
10,01 = 1
V TrEsp
4π0
j (0, 1)
i (1, 0)qB
qA
~ri − ~rj
(4)
NX
MX
i=1
j=1
Excitation energy transfer couplings for all three FMO
complexes calculated in the PDA and with the TrEsp
method are reported in Tab. S1.
Spectral density
We computed two point correlation functions using the
QM/MM first excited state energy trajectories obtained
for C. tepidum and for the ancestor. Subsequently we
fourier transformed and employed the harmonic approx-
imation (see Ref.
[12]) to obtain the spectral densities
J(ω). A spectral density was computed for each BChl
for both structures (see Fig. S4). No large difference
is seen. We also averaged the spectral densities over all
BChl's (see Fig. S5 left panel). The two average spectral
densities are almost identical, therefore most changes in
the dynamics will originate from changes in the system
hamiltonian.
To run the dynamics of the system using the HEOM
approach, we fitted the atomistic spectral densities to a
three peak Drude-Lorentz spectral density
νkλkω
k + (ω + Ωk)2 +
ν2
in Ref.
k=1
as done, e.g.
[36]. The fitted parameters are
given in Tab. S3. The fitted spectral densities are shown
on the right hand side panels of Fig. S5.
νkλkω
k + (ω − Ωk)2
ν2
(5)
,
(cid:20)
3X
J(ω) =
(cid:21)
13
Figure S1. Rooted phylogenetic tree for fmoA computed using MAFFT and TranslatorX [29] for the alignment and PHYML
[30]. Distances between branches correspond to difference in terms of sequence identity. Branches that are colored in light
green are branches where positive selection occurred.
14
Figure S2. Overlap of free energy variation histograms for all single-point mutations which could be applied to C. tepidum
(PDB: 3ENI, red) and P. aestuarii (PDB: 3EOJ, blue). Each bin contains the number of points with energy in that interval
normalized by the area under the entire histogram.
Table S1. Site couplings in C. tepidum, P. aestuarii and the ancestor in units of cm−1. Couplings were computed by using the
point-dipole approximation, with f = 1 according to Eq. 3, and also obtained from the transition charge using the electrostatic
potential (TrEsp) method, according to Eq. 4.
C. tepidum
P. aestuarii
Ancestor
-101.73
7.16
-6.27
7.65
-15.02
-9.38
-0.04
37.13
9.26
2.55
12.84
-3.66
1.41
Hnm Tresp [cm−1] Point dipole [cm−1] Tresp [cm−1] Point dipole [cm−1] Tresp [cm−1] Point dipole [cm−1]
1,2
1,3
1,4
1,5
1,6
1,7
1,8
2,3
2,4
2,5
2,6
2,7
2,8
3,4
3,5
3,6
3,7
3,8
4,5
4,6
4,7
4,8
5,6
5,7
5,8
6,7
6,8
7,8
-133.65
-9.53
-5.58
7.99
-17.05
5.31
0.13
-39.54
9.70
2.06
11.27
-3.52
1.57
124.56
-4.60
13.67
27.22
-2.98
-85.36
-26.06
42.09
-0.86
103.34
-2.96
7.25
-68.02
-9.36
9.59
-122.74
-5.53
-6.51
-9.44
16.27
-3.05
0.08
-37.17
7.60
-1.80
-7.25
2.60
2.00
80.99
-3.41
-12.23
7.73
-2.75
115.46
27.75
-49.82
-1.91
101.94
4.87
-7.71
-46.75
7.05
-10.76
-99.83
4.88
-6.87
8.54
-23.16
-8.37
-0.03
35.63
8.14
1.56
11.24
-11.89
1.98
-81.02
-5.48
-12.45
-2.21
2.69
-108.99
-2.26
-12.17
20.64
2.86
-102.29
-26.00
-66.42
-2.09
110.91
9.09
7.06
51.44
-5.36
-9.79
-102.28
-7.89
4.97
8.56
20.83
-11.33
0.38
-39.99
-9.69
2.11
-17.60
3.59
0.79
-125.38
1.80
-11.42
1.79
-2.89
103.72
-25.68
62.51
1.90
-86.77
18.89
7.45
-42.72
3.17
-10.38
-92.62
7.11
-6.49
9.17
-34.04
-15.35
0.81
37.48
10.92
1.79
18.87
3.60
0.88
-129.88
-3.91
-12.26
14.79
2.60
-95.63
-23.24
-62.86
-2.12
94.11
9.54
7.15
56.58
-2.33
-9.50
-103.75
-26.56
-67.90
-2.36
115.59
12.58
7.28
60.58
-5.57
-10.21
15
Figure S3. Histogram that summarizes the results of the computed ∆∆G for all single point mutations in P. aestuarii. For each
residue number (x-axis) we show a color that represents the free energy change associated with each of the 19 possible single
point mutations. The color is selected based on whether the free-energy change is stabilizing or not. Using σ = 0.46 kcal/mol, the
free energy changes where classified as neutral (green) for −σ ≤ ∆∆G ≤ σ, slightly destabilizing (yellow) for σ < ∆∆G ≤ 2σ,
destabilizing (orange) for 2σ < ∆∆G ≤ 3σ and highly destabilizing (burgundy) for 3σ < ∆∆G. Similarly, the stabilizing
mutations were classified as −2σ ≥ ∆∆G > −σ for slightly stabilizing (cyan), −3σ ≥ ∆∆G > −2σ for stabilizing (blue) and
∆∆G > −3σ for highly stabilizing (navy). On the top left-hand side we show the structure of the alpha-helix near BChl 8,
which can be modified in a stabilizing way, and on the right-hand side we highlight neighboring residues in another alpha helix
where mutations are more destabilizing.
16
Figure S4. The spectral density, J(ω), of each BChl in the harmonic approximation as a function of frequency. Each plot shows
a comparison of the spectral density for the C. tepidum FMO complex to that of the ancestral one. No large differences are
observed.
17
Figure S5. On the left hand side we show the average spectral density for the ancestor and the average spectral density for
the C. tepidum FMO complex. On the right hand side we show the Ohmic three-peak fit that is taken to obtain the spectral
densities used in the HEOM approach for the dynamics. The top panel shows the QM/MM spectral density in a solid line and
the Ohmic fit in a dashed line for the ancestral FMO complex. For the bottom panel the spectral densities are shown for C.
tepidum.
18
Table S2. Average distances and standard deviations of BChl-
protein distance. The distances where computed between Mg
atoms for each BChl and their neighboring protein environ-
ment. Distances where computed for BChls in C. tepidum, P.
aestuarii and the ancestor. For the protein environment we
selected the nitrogen atoms of HIS103, HIS290, HIS283 and
HIS138 for sites 1, 3, 4 and 6, the δ nitrogens of HIS289 for
site 7 and the peptide bond oxygens of TYR234 and TYR116
for sites 5 and 8. Residue indices refer to the C. tepidum se-
quences. Site 2 was not included as it does not bind to the
protein scaffold. Distances were computed and averaged over
the 40 ps production run.
Average BChl to protein distance hdBChl, P roteini
BChl C. tepidum [Å] P. aestuarii [Å] Ancestor [Å]
2.90 ± 0.26
3.89 ± 0.26
2.66 ± 0.22
2.94 ± 0.29
2.59 ± 0.18
3.65 ± 0.30
2.34 ± 0.16
2.82 ± 0.31
4.22 ± 0.37
2.58 ± 0.17
2.65 ± 0.21
2.94 ± 0.30
3.85 ± 0.66
2.43 ± 0.16
1
3
4
5
6
7
8
2.81 ± 0.22
4.31 ± 0.37
2.73 ± 0.18
2.47 ± 0.17
2.79 ± 0.24
3.82 ± 0.26
2.23 ± 0.13
Table S3. Drude-Lorentz parameters used to fit the spectral
density for the ancestor and for C. tepidum.
C. tepidum Ancestor
66.713
72.469
94.462
506.44
11.502
215.99
1229.0
5.1297
183.72
66.713
86.552
1.0000
809.03
13.142
199.53
583.76
1.6977
320.78
1/ν1[fs]
λ1[cm−1]
Ω1[cm−1]
1/ν2[fs]
λ2[cm−1]
Ω2[cm−1]
1/ν3[fs]
λ3[cm−1]
Ω3[cm−1]
19
coupled BChl molecules using QMaster [35, 58]. The
spectral densities were taken to be three peak fits of the
atomistic spectral density, as described previously. The
resulting population for the ancestor and C. tepidum in
shown in Fig. S8, we see that site 1 and 2 are less strongly
coupled for the ancestor than they are for C. tepidum.
One can see this also from the Hamiltonians as given in
Tab. S1. The coherence between site 1 and 2 is shown in
Fig. S9. Here we do not see a large difference between
the two species and this further confirmes the fact that
quantum coherence does not appear to have been selected
for.
Absorption, CD and LD spectra
In Fig. S6 we show the computed CD and LD spectra
for the ancestor and for C. tepidum at 300K. In Fig. S7
we show the comparison of our absorption, CD and LD
spectra computed using QM/MM results at 300K to the
experimental spectra of Ref. [63], for C. tepidum. We ob-
serve similar peaks in the absorption spectrum although
the 300K spectrum is much broader. The fact that the
spectrum is broader at 300K is expected because higher
temperatures lead to more motion and thus more noise
in the system. We also have a similar trend for the CD
and LD spectra as that seen experimentally.
Population dynamics
Using the HEOM approach [32–34] we computed the
time evolution of the density matrix of our system of 8
20
Figure S6. Comparison of simulated CD and LD spectra of the ancestral FMO complex (blue line) to that of current day C.
tepidum at 300K (grey dashed line) in arbitrary units. Spectra were computed using the Qy transitions obtained by using
QM/MM.
Figure S7. Comparison of our QM/MM spectra to the exper-
imental spectra for C. tepidum of Ref. [63].
21
Figure S8. Excited state population of each BChl as a func-
tion of time for an initial excited state in BChl 1 (dashed line,
C. tepidum, solid, Ancestor) at 300K.
Figure S9. Coherence between the excited states of BChls 1
and 2 as a function of time. These results were obtained by
propagating the density matrix of the system using HEOM at
300K and with BChl 1 as the initial state.
C. tepidumancestorC. tepidumancestor |
1310.6201 | 1 | 1310 | 2013-10-23T12:21:11 | Spectroscopic investigation of local mechanical impedance of living cells | [
"physics.bio-ph"
] | The mechanical properties of PC12 living cells have been studied at the nanoscale with a Force Feedback Microscope using two experimental approaches. Firstly, the local mechanical impedance of the cell membrane has been mapped simultaneously to the cell morphology at constant force. As the force of the interaction is gradually increased, we observed the appearance of the sub-membrane cytoskeleton. We shall compare the results obtained with this method with the measurement of other existing techniques. Secondly, a spectroscopic investigation has been performed varying the indentation of the tip in the cell membrane and consequently the force applied on it. In contrast with conventional dynamic atomic force microscopy techniques, here the small oscillation amplitude of the tip is not necessarily imposed at the cantilever first eigenmode. This allows the user to arbitrarily choose the excitation frequency in developing spectroscopic AFM techniques. The mechanical response of the PC12 cell membrane is found to be frequency dependent in the 1 kHz - 10 kHz range. The damping coefficient is reproducibly observed to decrease when the excitation frequency is increased. | physics.bio-ph | physics |
Spectroscopic investigation of local mechanical impedance of living cells
Luca Costa1,2, Mario S. Rodrigues3, N´uria Benseny-Cases4, V´eronique Mayeux1, Joel Chevrier5,6, Fabio Comin1
1) European Synchrotron Radiation Facility, 6 rue Jules Horowitz BP 220, 38043 Grenoble Cedex,
France
2) Universit´e Joseph Fourier BP 53, 38041 Grenoble Cedex 9, France
3) CFMC/Dep. Fisica, Faculdade de Ciencia, Universidade de Lisboa, Campo Grande, 1749-016
Lisboa, Portugal
4) Astbury Centre for Structural Molecular Biology, Leeds University, Leeds, UK
5) CNRS, Inst NEEL, F-38042 Grenoble, France
6) Universit´e Grenoble Alpes, Inst NEEL, F-38042 Grenoble, France
The mechanical properties of PC12 living cells have been studied at the nanoscale with a Force Feedback
Microscope using two experimental approaches. Firstly, the local mechanical impedance of the cell membrane
has been mapped simultaneously to the cell morphology at constant force. As the force of the interaction is
gradually increased, we observed the appearance of the sub-membrane cytoskeleton. We shall compare the
results obtained with this method with the measurement of other existing techniques. Secondly, a spectroscopic
investigation has been performed varying the indentation of the tip in the cell membrane and consequently
the force applied on it.
In contrast with conventional dynamic atomic force microscopy techniques, here the small oscillation amplitude
of the tip is not necessarily imposed at the cantilever first eigenmode. This allows the user to arbitrarily choose
the excitation frequency in developing spectroscopic AFM techniques. The mechanical response of the PC12
cell membrane is found to be frequency dependent in the 1 kHz - 10 kHz range. The damping coefficient is
reproducibly observed to decrease when the excitation frequency is increased.
September 4, 2018
1
Introduction:
Atomic Force Microscopes (AFMs) are intensively used for
studying cells. Because they provide the morphology of the
specimens at the nanoscale, AFMs have been employed for
imaging the cell surface and the submembrane cytoskele-
ton [1, 2, 3]. In addition, AFMs are nowadays extensively
used for molecular recognition experiments and to explore
the energy landscape of receptor-ligand interactions in liv-
ing cells [4, 5]. Another central aspect is the possibility to
apply a force on the cell membrane and measure the cell
elasticity. The mechanical response of the cells is a key ob-
servable for diseases diagnostics [6] and cell signaling [7, 8].
More generally, the elasticity is involved in many of the
physiological processes performed by the cell. Due to the
cell viscoelastic behavior [9, 10], the observed mechanical
properties could change significantly depending on the fre-
quency probed during an experiment. In this mainframe we
focus on the measurement of the mechanical impedance of
PC12 living cells employing atomic force microscopy meth-
ods.
In conventional static AFMs, the tip is slowly approached
close to the cell membrane and a force vs indentation curve
is then recorded. Depending on the geometry and the na-
ture of the mechanical contact between the tip and cell,
different contact models can be employed to extract the
intrinsic elasticity of the cells [11, 12]. The approach is
generally statistic since a set of approach curves needs to
be recorded to properly quantify the cell Young modulus.
A two-dimensional force map on the cell may be recorded to
get the Young modulus of each cell point [13, 14]. This tech-
nique is often time-consuming because a large numbers of
force curves needs to be acquired. Recently, advanced AFM
operational schemes have been proposed [15, 16] for mea-
suring nanomechanical properties of soft samples. These
methods, called multifrequency AFM, allow the user to si-
multaneously acquire the topography and the mechanical
properties of the specimens. This is possible either moni-
toring the higher harmonics excited while the tip is inter-
acting with the sample [16], either by exciting and moni-
toring the 2nd cantilever eigenmode [17]. As a consequence,
the sample elasticity is probed at frequencies linked to the
cantilever eigenmodes. These techniques are much faster
than the conventional force mapping.
We have introduced an alternative operational scheme
based on fiber optic detection system, called Force Feed-
back Microscope (FFM) [18, 19].
It allows the user to
simultaneously measure the static force, the elastic force
gradient and the damping coefficient, fully characterizing
the interaction between the AFM probe and the specimen.
The static force is the output of an active feedback loop
that controls the average position of the tip in the space.
A sub-nanometric oscillation amplitude is then imposed to
the tip for measuring the force gradient and the damping
coefficient. The use of small oscillation amplitudes indi-
cates the Force Feedback Microscope to be essentially a
linear AFM. At the solid/liquid interface soft cantilevers
with a stiffness in the order of 0.01 N/m are employed.
The use of soft cantilevers and small oscillation amplitude
is essential to reduce the invasiveness of the AFM experi-
ment [20] when soft samples have to be studied. A central
aspect of the Force Feedback Microscope is the capability
to use as feedback signal to record the sample morphol-
ogy either the force, either the force gradient, either the
1
damping coefficient. One out of these three different quan-
tities can be arbitrarily kept constant, providing a contrast
in the other two physical observables. In these conditions,
the contrast is intrinsically dependent on the choice of the
feedback signal, the magnitude of the setpoint, the nature
of the interaction and the local change of the interaction
on the specimens [19]. An additional key point of the Force
Feedback Microscope (FFM) is the possibility to arbitrarily
choose the excitation frequency imposed to the cantilever.
This capability is here exploited to characterize the me-
chanical response of PC12 living cells in the 1 kHz - 10
kHz frequency range. For this purpose, PC12 living cells
have been previously characterized in conventional static
mode. Subsequently, cells have been imaged at different
constant forces in the FFM mode. The topography is ac-
quired simultaneously to the measured sample stiffness and
the damping coefficient. Finally, a set of indentation curves
at different excitation frequencies has been recorded on the
top of the PC12 living cells.
2 Materials and methods:
Cell line
PC12 were obtained from ATCC. PC12 is a cell line derived
from a pheochromocytoma of the rat adrenal medulla and is
extensively used as a neuronal model. In the experimental
conditions used the cells were not differentiated.
Cell culture
Some existing protocols have been followed ([21, 22]). The
PC12 cells are cultivated inside 250 ml Flasks produced
by BD Falcon. The cell medium is a solution of Minimum
Essential Medium, with 10% Fetal Bovine Serum, supple-
mented with antibiotics 1% Penicillin-Streptomycin at 100
ng/ml. All the products are produced by Life technolo-
gies. The medium is changed every three days. Cells are
cultivated inside an incubator at 37◦C, 5 %CO2.
Glass support for AFM measurements
The glass coverslips (Thermanox, 13 mm diameter) are pre-
viously autoclave-ted for sterilization. The coverslips are
deposited inside the 4 wells Culture plates (Thermo Scien-
tific Nunc dishes IVF ), steryle. 500 µl of Fibronectin 1 µg
ml
(Life technologies) are deposited in each culture plate and
the sample is then left incubated for 60 minutes at room
temperature. The Fibronectin is then removed and each
culture plate is washed 4 times with sterile PBS.
500 µl of MEM with 10% Fetal Bovine Serum and 1%
Penicillin-Streptomycin are deposited in each culture plate.
Cells transfer on the glass support for the
AFM measurements:
The buffer inside the flask is removed. A solution of 15
ml PBS, 0.25% Trypsin at 1 mg/ml is transferred inside
the flask. The flask is then left incubated for 30 minutes
at 37◦C. The cells are now suspended in the solution. 500
µl of fetal bovine serum are added to the solution. The
solution is centrifuged for 5 minutes at 12500 rpm inside a
15 ml centrifuge tube. The solution is removed leaving the
cells at the bottom of the tube. 2 ml of MEM, 10% fetal
bovine serum, 1% Penicillin-Streptomycin at 100 ng/ml are
inserted in the tube. Cells are mechanically re-suspended
using a Pasteur pipette. The cells concentration is then
calculated using a Neubauer cell. Cells have been usually
deposited in each culture plate on the top of the coverslip
glass at a concentration of 1 million/ml. The cells are then
left in the incubator at 37◦C, 5 %CO2 for one/two nights
and then measured with the FFM. The imaging buffer is
MEM, 1% Penicillin-Streptomycin at 100 ng/ml.
Figure 1: Spectroscopy on the glass in liquid buffer. Force
gradient (a) and damping factor γ (b) as a function of the
excitation frequency. Excitation frequencies: Blue = 1.11
kHz, Red = 3.11 kHz, Green = 9.11 kHz, Black = 11.11
kHz.
Force Feedback Microscopy
The Force Feedback Microscope is a custom - made AFM
based on a fiber optic interferometer detection system [18].
Triangular and rectangular MLCT cantilevers provided by
Bruker with nominal spring constants of 0.01 N/m and
0.02 N/m respectively have been employed. Both the can-
tilevers have a nominal tip radius of 20 nm.
An Asylum Research fluid cell has been modified to fit prop-
erly in the home made Force feedback Microscope. The
glass coverslips have been fixed inside the liquid cell. Cells
have been kept at 37◦C controlling the temperature of the
2
Figure 2: PC12 living cells imaged in conventional contact
mode. Scan area = 90 x 90 µm2. Several cells of different
shapes are here imaged.
imaging buffer during the measurements. A central aspect
of the Force Feedback Microscope is the calibration of the
instrument. For this purpose, a set of approach curves
between the tip and the glass has been recorded for each
excitation frequency ω.
∇F = a [cos(φ∞) − n cos(φ)]
γ =
a
ω
[sin(φ∞) − n sin(φ)]
(1)
(2)
Equation (1) and (2) convert the measured values n and φ
into the interaction properties ∇F and γ. φ is the phase
between the oscillation of the tip and excitation at the can-
tilever base, whereas n is the so-called normalized ampli-
tude. Equation (1) and (2) can be used mainly because
the oscillation amplitude imposed to the tip is kept below
1 nm. It follows that the force feedback microscope can be
considered as a linear AFM.
In this case, assuming the integrated force gradient to be
equal to the measurement of the static force between the
AFM probe and the glass, the key point is to determine
the parameters φ∞ and a. The protocol of calibration is
explained in details elsewhere [18].
In figure 1 the elas-
tic force gradient and the damping coefficient measured on
the glass at 4 different excitation frequencies are reported.
The elastic force gradient is found to be independent of the
frequency probed (figure 1a). The damping coefficient is
observed to decrease at larger excitation frequencies (fig-
ure 1b). Since the dissipative force is proportional to the
damping coefficient times the speed of the tip, we do not ob-
serve a significant decrease in the force needed to maintain
the tip oscillations constant when the excitation frequency
is increased. In other words, the dissipative force exerted
by the glass on the AFM tip is found to be constant in the
1 kHz - 10 kHz range.
3
Figure 3: Indentation static curve on a PC12 (blue) fit-
ted with Hertz model(red). The tip is modeled as a four-
sided pyramidal indenter. The Young's modulus extracted
is 1716 Pa using a Poisson ratio equal to 0.5 and θ = 25◦.
3 Results:
Characterization in conventional
mode
static
Cells have been firstly characterized with conventional
AFM static mode.
In figure 2 a typical image acquired
at a constant repulsive force of 100 pN is reported. Since
in our custom - made AFM it is not possible to laterally
offset the sample from the tip, a scan area of 100 x 100
µm2 is the only sample surface accessible to obtain an im-
age. As a consequence, the cell concentration employed for
the cell culture has been refined in order to obtain a cell
density on the glass support similar to the one presented in
the image. Indeed, the presence of the glass support in the
scan area is a compulsory condition to calibrate the FFM.
Once the cells are successfully imaged, a set of inden-
tation force curves has been acquired on the living cells.
The tip can be modeled as a four-sided pyramidal inden-
ter. In this case, a relationship between the force F and
the indentation δ [23] can be given by
F =
3Etanθ
4(1 − ν 2)
δ2
(3)
where E is the Young modulus and ν is the Poisson ratio of
the cell. ν is assumed to be equal to 0.5. The nominal θ of
the probes supplied with the MLCT cantilevers is 25◦. In
figure 3 a typical force vs indentation curve is shown. The
measured Young modulus is 1716 Pa which is comparable
to the one measured on PC12 living cells in previous works
[24].
Characterization at constant force in force
feedback mode
Cells have been imaged in Force Feedback Microscopy mode
at a constant repulsive force of 500 pN. In this imaging
mode, the force is kept constant during the scan and the
topography is acquired simultaneously to the force gradient
and the damping factor. An amplitude of 0.4 nm has been
imposed to the tip at 2.2 kHz. The Force Feedback Micro-
scope has been then calibrated [18] with a set of approach
effects can be due to the conformation of the cytoskeleton,
in particular of the actin filaments and the microtubules in-
side the cell as suggested by [14, 25, 26]. These effects can
be also due to the higher influence of the glass substrate
when the measurements of the stiffness and the damping
factor are carried out at the periphery of the cells. The
measured stiffness of the cells is comparable to the stiff-
ness measured in multi-harmonic atomic force microscopy
on different cells [16].
In a different experiment, PC12 cells have been imaged
at the constant force of 1 nN. An amplitude of 0.4 nm
has been imposed to the tip at 7.78 kHz.
In figure 4e
is reported the topography, in figure 4f the force (error
signal),
in figure 4g the elasticity and in figure 4h the
damping coefficient.
Figure
fila-
ments/microtubules which are however hardly measured in
the elasticity and damping coefficient images. In analogy
with the measurement at constant force of 500 pN, stiffness
and damping factor are observed to be lower in the center
of the cells than at the periphery of the cells.
presence
actin
of
4e
reveals
the
Characterization at constant force and tun-
able frequency
The capability of the FFM to arbitrarily choose the excita-
tion frequency of the AFM tip allows the user to measure
the sample force gradient and the sample damping factor
at the desirable frequency. For this purpose a set of cali-
bration curves on the top of the glass substrate have been
performed at 2.25 kHz and then at 13.25 kHz. The oscil-
lation amplitude imposed to the tip is 0.4 nm. Once the
FFM is calibrated, one image at constant repulsive force of
50 pN has been acquired at the excitation frequency of 2.25
kHz (figure 5a,b,c,d). Consequently, the same scan area has
been imaged a second time at constant repulsive force of 50
pN, imposing an excitation frequency of 13.25 kHz (figure
5e,f,g,h).
In analogy with the measurement presented in
figure 4, the cell center is again observed to be softer than
the cell borders.
The stiffness and the damping factor measured at 2.25
kHz are nosier than those measured at 13.25 kHz mainly
because of the larger influence of the noise 1
f . In order to
run an acceptable measurement at low frequency, the scan
speed at 2.25 kHz has been set twice slower than the one
at 13.25 kHz.
The changes in the sample morphology between figure
5a and figure 5e are due to the fact that the cells are alive.
We observe consistent differences between the images of the
elasticity and the damping factor. The cells are found to be
stiffer at 13.25 kHz than at 2.25 kHz by a factor five. The
damping factor is instead observed to be larger at 2.25 kHz
than at 13.25 kHz. The decrease of the damping factor once
the excitation frequency is increased is a behavior observed
on the spectroscopy performed on the glass (figure 1).
Indentation curves
A spectroscopic investigation of the PC12 cells has been de-
veloped through the acquisition of tip-cell approach curves.
The protocol is the following:
Figure 4: FFM images of PC12 living cells acquired at
a constant force of 500 pN and 1 nN. In the images at
500 pN constant force (a,b,c,d) the scale bar is 30 µm. a)
Sample topography, b) force (error), c) force gradient and
d) damping factor.
In the images at 1 nN (e,f,g,h) the scale bar is 90 µm. e)
Sample topography, f) force (error), g) force gradient and
h) damping factor.
curves performed on the glass. Subsequently, the tip has
been brought in contact with the sample at a constant re-
pulsive force of 500 pN and the image has been acquired.
In figure 4a is reported the topography and in figure 4b
the force between the tip and the sample which is the er-
ror signal. In addition, 4c is the measurement of the force
gradient, whereas 4d is the damping factor.
The membrane is found to be softer in the center of the
cell than at the periphery. In analogy with the stiffness,
the damping factor is observed to be lower in the center of
the cell than at the borders of the cell. In fibroblasts, these
4
• The cells were imaged in conventional contact mode.
• The Force Feedback Microscope was calibrated with
a set of approach curves at different frequencies on
the glass. A set of parameters (φinf ty)i and ai for
each excitation frequency ωi were obtained. The force
gradients ∇Fi are equal for each excitation frequency
as shown in figure 1a.
• A set of approach curves at different frequency on the
PC12 was performed in the highest point of the cell
topography.
For an applied force of 500 pN, the observed cell elasticity in
all the measurements has always been found in the order of
0.01 N/m. An example of such an experiment is presented
in figure 6 We notice that the damping rate is decreasing
Figure 5: FFM images of PC12 living cells acquired at a
constant force of 50 pN. Scan area = 90 x 90 µm2. a)
Sample topography, b) force (error), c) force gradient and
d) damping factor at 2.25 kHz. e) Sample topography, f)
force (error), g) force gradient and h) damping factor at
13.25 kHz
Figure 6: Spectroscopy of the PC12 in liquid buffer. Force
gradient (a), damping factor (b) and dissipative force (in-
set) as a function of the excitation frequency. Excitation
frequencies: Blue = 1.13 kHz, Red = 5.13 kHz, Green =
7.13 kHz, Black = 11.13 kHz.
when the excitation frequency increases, whereas the force
gradient is increasing with the frequency. The increase of
the elasticity with the frequency has not been observed
to be reproducible in all the measurements performed. In
particular, we observe a decrease of the cell elasticity at fre-
quencies larger than 13 kHz. The decrease of the damping
factor with the frequency is, on the contrary, reproducible.
The resulting dissipative force, presented in the inset of
figure 6b, is observed to be poorly dependent on the fre-
quency. According to [10], the damping factor is expected
to increase with the excitation frequency in the 1 Hz - 100
5
Hz range. However, in our case the oscillation amplitude
imposed to the tip is 2 orders of magnitude lower than the
usual amplitudes used for studying living cells. Since in
the present case the oscillation amplitude is much smaller
than the cell thickness, we think that the effect of of what
is below the membrane may be smaller when compared to
the case were the amplitude of oscillation is larger than to
the cell thickness. In addition, a small oscillation ampli-
tude may induce measurements of stiffness and damping
factor which are more localized on the cell membrane than
in an experiment where larger oscillation amplitudes are
imposed to the tip.
Analysis of elastic properties in dynamic
mode:
Focusing on the elasticity, from equation (3) we can define
the force gradient as a function of the tip indentation as
measurements performed has always been found in the or-
der of ten(s) of kPa in this frequency range. The measured
Young modula are therefore observed to be between one
and two orders of magnitude larger than the static values
(figure 3).
Spatial variation analysis of elastic proper-
ties:
The spatial variation of the cell elasticity can be studied
with the FFM during the acquisition of the cell morphology
as presented in figure 4 and figure 5. In analogy with the
study of the cell elasticity in dynamic mode, it is possible
to extract the spatial variation of the Young modulus. The
simultaneous measurement of the elasticity and the applied
force allows us to evaluate the Young modulus of the cells
through the relation 5.
∇F =
3Etanθ
2(1 − ν 2)
δ
(4)
(∇F )2
F
≈
3Etan(θ)
1 − ν 2
(5)
In particular, the data presented in figure 6a have been
used to extract the cell Young modulus as a function of the
excitation frequency. The Young modula are found to be
equal to 8690 Pa at 1.13 kHz, 18604 Pa at 5.13 kHz, 26385
Pa at 7.13 kHz and 27023 Pa at 11.13 kHz. The data are
reported in figure 7.
In figure 8 the extracted Young modulus of PC12 liv-
ing cells imaged at a constant repulsive force of 1 nN is
reported.
Figure 7: Analysis of the PC12 elastic properties through
indentations curves. a) Force gradient as a function of the
tip indentation. a) ω = 1.13 kHz, b) ω = 5.13 kHz, c) ω
= 7.13 kHz, d) ω = 11.13 kHz. The lines in orange are the
experimental linear fit of the cell elasticity using equation
(4).
In the last decades, the Force Modulation technique [27]
has been applied to the study of the variation of the elas-
ticity of the cells, typically at frequencies lower than 100
Hz [10, 28]. In this frequency range the studied cells show
an increase of the young modulus with frequency following
specific power laws.
Here the study has been extended to frequencies larger than
than 1 kHz. As reported in the previous section, no clear
reproducibility of the increase of the cell stiffness has been
measured. However, the observed cell elasticity in all the
Figure 8: Spatial variation of the Young modulus of PC12
living cells imaged at constant repulsive force of 1 nN and at
the excitation frequency of 7.78 kHz. The Young modulus
is extracted using relation (5).
The scale-bar of figure 8 is limited to 3 MPa to enhance
the contrast in the spatial variation of the cell Young mod-
ulus. The periphery of the cells is observed to be much
stiffer than the center, reaching values larger than 1 MPa.
The Young modulus in the center of the cell is observed
to vary from 50 kPa up to 400 kPa. The extracted Young
modulus may be compared with the values estimated on
different living cells in multi-harmonics AFM [16], reveal-
ing a consistent agreement.
6
4 Conclusion:
Concluding, the Force Feedback Microscope measures si-
multaneously the force, the force gradient and the damping
factor at different excitation frequencies, fully characteriz-
ing the interaction between the AFM tip and the sample.
We developed two spectroscopic protocols to study living
cells with a Force Feedback Microscope. The techniques
have been applied to the study of PC12 living cells. At
first, cells have been characterized with images at differ-
ent constant forces and at different excitation frequencies.
Then, the response of the cell membrane has been charac-
terized through the acquisition of approach curves at differ-
ent frequencies. The local mechanical impedance, that is
the elasticity and the damping factor, have been measured
and characterized in the 1 kHz - 10 kHz range. Our mea-
surements have been compared with data existing in liter-
ature revealing that the amplitude of oscillation applied on
the tip is a crucial parameter when the cell elasticity and
dissipation have to be characterized. The local mechani-
cal impedance of the PC12 living cell has been character-
ized in the x/y plane simultaneously to the cell morphology
in analogy with already existing technique such as multi-
harmonics AFM [16].
Finally, it is clear that one of the future objectives is the
extension of the lower and upper limits of the available
excitation frequencies, 1 kHz and 15 kHz respectively.
ACKNOWLEDGMENTS
Luca Costa acknowledges COST Action TD 1002. Mario S.
Rodrigues acknowledges financial support from Funda¸cao
para a Ciencia e Tecnologia SFRH/BPD/69201/2010. This
work was performed at the Surface Science Laboratory of
the ESRF and at the PSB laboratories in the Carl-Ivar
Brand´en building in Grenoble.
References
[1] F. Braet, C. Rotsch, E. Wisse, and M. Radmacher,
"Comparison of fixed and living liver endothelial
cells by atomic force microscopy," Applied Physics A,
vol. 66, no. 1, pp. S575 -- S578, 1998.
[2] S. Kasas and A. Ikai, "A method for anchoring round
shaped cells for atomic force microscope imaging,"
Biophysical Journal, vol. 68, no. 5, pp. 1678 -- 1680,
1995.
[3] C. L. Grimellec, E. Lesniewska, M.-C. Giocondi,
E. Finot, V. Vi, and J.-P. Goudonnet, "Imaging of
the surface of living cells by low-force contact-mode
atomic force microscopy," Biophysical Journal, vol. 75,
no. 2, pp. 695 -- 703, 1998.
[4] D. J. Muller and Y. F. Dufrene, "Atomic force mi-
croscopy as a multifunctional molecular toolbox in
nanobiotechnology," nature nanotechnology, vol. 3,
pp. 261 -- 269, 2008.
[5] D. J. Muller, J. Helenius, D. Alsteens, and Y. F.
Dufrene, "Force probing surfaces of living cells to
molecular resolution," Nature chemical biology, vol. 5,
no. 6, pp. 383 -- 390, 2009.
[6] M. Plodinec, M. Loparic, C. A. Monnier, E. C. Ober-
mann, R. Zanetti-Dallenbach, P. Oertle, J. T. Hyotyla,
U. Aebi, M. Bentires-Alj, L. Y. H., and C.-A. Schoe-
nenberger, "The nanomechanical signature of breast
cancer," Nat Nano, vol. 7, pp. 757 -- 765, Nov. 2012.
[7] A. J. Engler, S. Sen, H. L. Sweeney, and D. E. Discher,
"Matrix elasticity directs stem cell lineage specifica-
tion," Cell, vol. 126, no. 4, pp. 677 -- 689, 2006.
[8] A. E. Brown, A. Hategan, D. Safer, Y. E. Gold-
man, and D. E. Discher, "Cross-correlated tirf/afm
reveals asymmetric distribution of
force-generating
heads along self-assembled, synthetic myosin fila-
ments," Biophysical Journal, vol. 96, no. 5, pp. 1952 --
1960, 2009.
[9] B. Fabry, G. N. Maksym, J. P. Butler, M. Glogauer,
D. Navajas, and J. J. Fredberg, "Scaling the mi-
crorheology of living cells," Phys. Rev. Lett., vol. 87,
p. 148102, Sep 2001.
[10] J. Alcaraz, L. Buscemi, M. Grabulosa, X. Trepat,
B. Fabry, R. Farr, and D. Navajas, "Microrheology
of human lung epithelial cells measured by atomic
force microscopy," Biophysical Journal, vol. 84, no. 3,
pp. 2071 -- 2079, 2003.
[11] F. Rico, P. Roca-Cusachs, N. Gavara, R. Farre,
M. Rotger, and D. Navajas, "Probing mechanical
properties of living cells by atomic force microscopy
with blunted pyramidal cantilever tips," Phys. Rev.
E, vol. 72, p. 021914, Aug 2005.
[12] F. Gaboriaud and Y. F. Dufrne, "Atomic force mi-
croscopy of microbial cells: Application to nanome-
chanical properties,
surface forces and molecular
recognition forces," Colloids and Surfaces B: Bioin-
terfaces, vol. 54, no. 1, pp. 10 -- 19, 2007.
[13] U. G. Hofmann, C. Rotsch, W. J. Parak, and M. Rad-
macher, "Investigating the cytoskeleton of chicken car-
diocytes with the atomic force microscope," Journal of
Structural Biology, vol. 119, no. 2, pp. 84 -- 91, 1997.
[14] H. Haga, S. Sasaki, K. Kawabata, E. Ito, T. Ushiki,
and T. Sambongi, "Elasticity mapping of living fibrob-
lasts by afm and immunofluorescence observation of
the cytoskeleton," Ultramicroscopy, vol. 82, no. 14,
pp. 253 -- 258, 2000.
[15] R. Garcia and E. T. Herruzo, "The emergence of
multifrequency force microscopy," Nat Nano, vol. 7,
pp. 217 -- 226, Apr. 2012.
[16] A. Raman, S. Trigueros, A. Cartagena, A. P. Z.
Stevenson, M. Susilo, E. Nauman, and S. A. Contera,
"Mapping nanomechanical properties of live cells us-
ing multi-harmonic atomic force microscopy," Nature
Nanotechnology, vol. 6, pp. 809 -- 814, Dec. 2011.
7
[17] D. Martinez-Martin, E. T. Herruzo, C. Dietz,
J. Gomez-Herrero, and R. Garcia, "Noninvasive pro-
tein structural flexibility mapping by bimodal dy-
namic force microscopy," Phys. Rev. Lett., vol. 106,
p. 198101, May 2011.
[18] M. S. Rodrigues, L. Costa, J. Chevrier, and F. Comin,
"Why do atomic force microscopy force curves still
exhibit jump to contact?," Applied Physics Letters,
vol. 101, no. 20, p. 203105, 2012.
[19] L. Costa, M. S. Rodrigues, E. Newman, C. Zubieta,
J. Chevrier, and F. Comin, "Imaging material prop-
erties of biological samples with a force feedback mi-
croscope," Journal of Molecular Recognition, vol. ac-
cepted, 2013.
[20] H. V. Guzman, A. P. Perrino, and R. Garcia, "Peak
forces in high-resolution imaging of soft matter in liq-
uid," ACS Nano, vol. 7, no. 4, pp. 3198 -- 3204, 2013.
[21] C. Viro, I. Mchaly, H. Aptel, S. Puech, J. Valmier,
F. Bancel, and G. Dayanithi, "Rapid inhibition of
ca2+ influx by neurosteroids in murine embryonic sen-
sory neurones," Cell Calcium, vol. 40, no. 4, pp. 383 --
391, 2006.
[22] S. Andr´e, H. Boukhaddaoui, B. Campo, M. Al-
Jumaily, V. Mayeux, D. Greuet, J. Valmier, and
F. Scamps, "Axotomy-induced expression of calcium-
activated chloride current in subpopulations of mouse
dorsal root ganglion neurons," Journal of Neurophys-
iology, vol. 90, no. 6, pp. 3764 -- 3773, 2003.
[23] G. Bilodeau, "Regular pyramid punch problem," Jour-
nal of applied mechanics, vol. 59, p. 519, 1992.
[24] C.-T. Chang, C.-C. K. Lin, and M. S. Ju, "morphology
and ultrastructure-related local mechanical properties
of pc-12 cells studied by integrating atomic force mi-
croscopy and immunofluorescence imaging," Journal
of Mechanics in Medicine and Biology, vol. 12, no. 05,
p. 1250032, 2012.
[25] C. Rotsch and M. Radmacher, "Drug-induced changes
of cytoskeletal structure and mechanics in fibroblasts:
An atomic force microscopy study," Biophysical Jour-
nal, vol. 78, no. 1, pp. 520 -- 535, 2000.
[26] J. Li, A. Shariff, M. Wiking, E. Lundberg, G. K.
Rohde, and R. F. Murphy, "Estimating microtubule
distributions from 2d immunofluorescence microscopy
images reveals differences among human cultured cell
lines," PLoS ONE, vol. 7, p. e50292, 11 2012.
[27] M. Radmacher, R. Tillmann, and H. Gaub, "Imaging
viscoelasticity by force modulation with the atomic
force microscope," Biophysical Journal, vol. 64, no. 3,
pp. 735 -- 742, 1993.
[28] T. Luque, E. Melo, E. Garreta, J. Cortiella, J. Nichols,
R. Farr, and D. Navajas, "Local micromechanical
properties of decellularized lung scaffolds measured
with atomic force microscopy," Acta Biomaterialia,
vol. 9, no. 6, pp. 6852 -- 6859, 2013.
8
|
1901.03066 | 1 | 1901 | 2019-01-10T09:09:34 | Pearl-Necklace-Like Local Ordering Drives Polypeptide Collapse | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.BM"
] | Collapse of the polypeptide backbone is an integral part of protein folding. Using polyglycine as a probe, we explore the nonequilibrium pathways of protein collapse in water. We find that the collapse depends on the competition between hydration effects and intra-peptide interactions. Once intra-peptide van der Waal interactions dominate, the chain collapses along a nonequilibrium pathway characterized by formation of pearl-necklace-like local clusters as intermediates that eventually coagulate into a single globule. By describing this coarsening through the contact probability as a function of distance along the chain, we extract a time-dependent length scale that grows in linear fashion. The collapse dynamics is characterized by a dynamical critical exponent $z=0.5$ that is much smaller than the values of $z=1-2$ reported for non-biological polymers. This difference in the exponents is explained by the instantaneous formation of intra-chain hydrogen bonds and local ordering that may be correlated with the observed fast folding times in proteins. | physics.bio-ph | physics | Pearl-Necklace-Like Local Ordering Drives Polypeptide Collapse
Suman Majumder,1 Ulrich H. E. Hansmann,2 and Wolfhard Janke1
1Institut fur Theoretische Physik, Universitat Leipzig, Postfach 100 920, 04009 Leipzig, Germany
2Department of Chemistry and Biochemistry, University of Oklahoma, Norman, Oklahoma 73019, USA
(Dated: January 11, 2019)
Collapse of the polypeptide backbone is an integral part of protein folding. Using
polyglycine as a probe, we explore the nonequilibrium pathways of protein collapse in
water. We find that the collapse depends on the competition between hydration effects
and intra-peptide interactions. Once intra-peptide van der Waal interactions dominate,
the chain collapses along a nonequilibrium pathway characterized by formation of pearl-
necklace-like local clusters as intermediates that eventually coagulate into a single
globule. By describing this coarsening through the contact probability as a function of
distance along the chain, we extract a time-dependent length scale that grows in linear
fashion. The collapse dynamics is characterized by a dynamical critical exponent z = 0.5
that is much smaller than the values of z = 1 − 2 reported for non-biological polymers.
This difference in the exponents is explained by the instantaneous formation of intra-
chain hydrogen bonds and local ordering that may be correlated with the observed
fast folding times in proteins.
9
1
0
2
n
a
J
0
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
6
6
0
3
0
.
1
0
9
1
:
v
i
X
r
a
2
Changing the solvent condition from good to poor renders an extended polymer to undergo a collapse transition
by forming a compact globule [1, 2]. Both experiments [3, 4] and simulations [5, 6] indicate that a protein also
experiences such a collapse transition while folding into its native state. However, the nonequilibrium dynamics of the
collapse of proteins is only poorly understood and an active research topic [7]. Most previous studies consider only the
hydrophobicity of apolar side chains of amino acids in a protein as driving force for collapse [8, 9]. In the present paper
we focus instead on the contributions by intra-peptide interactions, present even for residues with no hydrophobic
or only weakly hydrophobic side chains [10 -- 13] where the collapse-driving forces are not necessarily proportional to
the exposed surface. Our test system is polyglycine, and has been chosen to connect our work with recent studies
of homopolymer collapse dynamics [14 -- 17] that found nonequilibrium scaling laws as known for generic coarsening
phenomena [18]. Our hope is to establish such scaling laws also for the collapse of proteins. As a first stride towards
this goal, here, we explore the kinetics of collapse of polyglycine.
Collapse of homopolymers was first described by de Gennes' seminal "sausage" model [19], but today the phe-
nomenological "pearl-necklace" picture by Halperin and Goldbart [20] is more commonly used, both for flexible
[14, 16, 17, 21 -- 25] and semiflexible polymer models [26, 27]. In this picture the collapse begins with nucleation of
small local clusters (of monomers) leading to formation of an interconnected chain of (pseudo-)stable clusters, i.e.,
the "pearl-necklace" intermediate. These clusters grow by eating up the un-clustered monomers from the chain and
subsequently coalesce, leading eventually to a single cluster. Finally, monomers within this final cluster rearrange to
form a compact globule.
Of central interest in this context is the scaling of the collapse time τc with the degree of polymerization N (the
number of monomers). While scaling of the form τc ∼ N z, where z is the dynamic exponent, has been firmly
established, there is no consensus on the value of z. Molecular dynamics (MD) simulations provide much smaller
values (z ≈ 1) than Monte Carlo (MC) simulations (z ≈ 2). This difference is often explained with the presence of
hydrodynamics in the MD simulations, but a value z ≈ 1 has been reported recently also for MC simulations [16].
The "pearl-necklace" stage or the cluster-growth kinetics can be understood by monitoring the time (t) dependence of
the mean cluster size Cs(t), the relevant length scale. By drawing analogy with coarsening ferromagnets, it has been
shown that scaling of the form Cs(t) ∼ tαc with growth exponent αc ≈ 1 holds for flexible homopolymers [14, 16].
Protein collapse is much less understood. While it has been shown by modeling a protein as semiflexible het-
eropolymer that the equilibrium scaling of the radius of gyration Rg with N is random-coil-like in a good solvent
and globule-like in a poor solvent [28, 29], there have been few attempts to explore nonequilibrium collapse pathways
3
FIG. 1. Time evolution of a short polypeptide. a) The upper panel shows snapshots from the time evolution for collapse
of (Gly)20 chain in water at Tq = 290 K, starting from an extended state at t = 0 ns. The lower panel shows the corresponding
residue contact maps where two residues along the chain are in contact if the distance between them is less than 1.5 nm. (b)
Time dependence of the squared radius of gyration R2
g(t) from five different runs. (c) Illustration of the structural evolution
of the chain during the collapse shown via structure factor S(q) as function of the wave vector q at the times presented in (a).
The dashed lines with power-law decay exponent 5/3 and 4 correspond to the expected behavior for an extended chain and
crumpled globule, respectively.
[30, 31], and the corresponding scaling laws are not known. In order to probe the existence of such nonequilibrium
scaling laws in protein collapse, we have simulated polyglycine chains (Gly)N of various numbers N of residues. This
choice allows us to probe in a systematic way the collapse of polypeptide chain, considering only homopolymers built
from the simplest amino acid, glycine. Our results show that in water there is a tug of war between collapse-disfavoring
hydration effects and collapse-favoring intra-peptide interactions. For longer chains (N ≥ 15) the intra-peptide in-
teractions win over the hydration effect leading to a collapse, making water in practice a poor solvent. We use these
longer polyglycine chains to shed light on the collapse kinetics, with an emphasis on the presence of nonequilibrium
scaling laws. Our results from all-atom MD simulations in the NVT ensemble using a hydrodynamics preserving
thermostat (see the Method section for details), suggest a collapse mechanism that relies on fast local ordering by
formation of pearl-necklace structures which eventually merge into a single globule. This process is characterized by
a dynamic critical exponent z = 0.5 much smaller than the exponents z = 1− 2 observed for non-biological polymers,
and we speculate that this quicker local ordering and collapse enables the fast folding times seen in proteins.
We begin our analysis with a rather short chain, i.e., (Gly)20. The time evolution snapshots during the collapse in
water at a temperature Tq = 290K, well below the corresponding collapse transition temperature, are shown in Fig.
1(a). In a protein, collapse leads eventually to folding characterized by formation of distinct native contacts among the
1101000.01100 0 ns 2 ns 5 ns 10 ns 0246810012run 1run 2run 3run 4run 5S(q)t [ns]Rg2(t) [nm2](b)q [nm-1]q-4q-5/3(c)2.54
residues. We show for this reason in the lower panel the residue contact maps where we define two residues as being in
contact if they are within a distance of rc = 1.5 nm. The red stripe along the diagonals depicts the self-contacts. The
size of the extended (Gly)20 chain is ≈ 2.0 nm, thus almost all the mutual distances between the residues fall under rc.
This makes it difficult to capture segregation or formation of any local structures on length scales comparable to rc.
Only late in the trajectories do we find a signature for loop formation, which is also apparent in the snapshot at t = 10
ns. Emergence of such loop is due a competition between the hydration effects and the intra-peptide interactions
leading to residue-residue contacts along the chain, although there are trapped water molecules. The interplay can
be deduced from the non-monotonous behavior of the squared radius of gyration R2
g as function of time in Fig. 1(b),
obtained from 5 independent runs. Note that for all the cases R2
g decays eventually to the equilibrium value.
In order to probe further the structural evolution of the chain along the collapse of (Gly)20, we calculate the static
structure factor S(q) at different times. Fig. 1(c) shows S(q) for the times corresponding to the snapshots. At t = 0
ns, within the range q ∈ [3, 30] nm−1, the chain can be described as an extended coil with S(q) ∼ q−1/ν [32], where
ν = 3/5 is the critical exponent describing the scaling of Rg ∼ N ν for a self-avoiding polymer. With time the decay
exponent should increase from −5/3 and is expected to approach −4, in order to be consistent with the globule-like
behavior of S(q) ∼ q−4 [32]. Although the slope in our data in Fig. 1(c) gradually increases with time, it does not
appear to approach −4. This again could be due to the still ongoing interplay between the hydration effect and the
intra-peptide interactions which hinders the chain to form a compact globule, however, extending the simulations up
to 20 ns does not change the overall behavior. Similar observations are made for all systems (Gly)N having a chain
length of N < 50 residue units.
For longer chains, the collapse is more pronounced, and we finally encounter characteristic features reminiscent of
the homopolymer collapse. For instance, in Fig. 2(a), we present snapshots of the collapse of (Gly)200 at Tq = 290
K. The sequence of these snapshots demonstrates a process that starts with local ordering of the residues along the
chain. These local structures later merge with each other before finally forming a single globule at t ≈ 20 ns. The
emergence of these local arrangements is similar to the formation of local clusters in the "pearl-necklace" picture of
homopolymer collapse [14, 16, 20, 21]. The resemblance becomes even more obvious when looking at the corresponding
contact maps in the lower panel. The box-like clustering along the diagonal indicates formation of "pearls" along
the chain (see particularly at t = 2 and 5 ns) that are reminiscent to the ones observed during the collapse of
semiflexible homopolymer in Ref. [27]. However, we do not see the anti-parallel hairpins that were associated with
this diamond-shaped internal orders within these boxes.
5
FIG. 2. Pearl-necklace formation during collapse of longer chains. (a) Same as in Fig. 1(a) but for (Gly)200 and
correspondingly at different times, as mentioned. (b) Time dependence of the squared radius of gyration R2
g(t) from 5 different
runs. (c) The structure factors S(q) at times that are presented in (a). There dashed lines have the same meaning as in Fig. 1
(c).
In order to check for the presence of a competition between hydration effects and the intra-peptide interactions
we probe again the time dependence of R2
g as measured in five independent runs. Data are presented in Fig. 2(b).
Unlike for the shorter (Gly)20 chain, the radius of gyration is now monotonically decreasing. This can be explained by
the assumption that for longer chains the intra-chain interactions overcome the hydration effects. A similar picture
emerges from Fig. 2(c). The plots of the structure factor S(q) as function of time demonstrate how the extended coil
behavior of S(q) ∼ q−5/3 at t = 0 ns gradually changes to a globule-like behavior of S(q) ∼ q−4 at t = 20 ns.
Next, we analyze the number of intra-molecular, i.e., protein-protein npp and inter-molecular, i.e., protein-water
npw hydrogen (H)-bonds. These quantities, measured for different N and normalized by the respective values at tf
(the maximum time up to which the simulations are run; for details see the Method section) are plotted as function
of time in the main frame of Fig. 3(a) and (b). Data for all N in Fig. 3(a) attain the value ≈ 1 at the same time,
demonstrating a reasonable overlap of the normalized data. Similarly happens in (b) the decay of npw to the saturation
value at almost the same time for different N , leading again to nicely overlapping curves. In the inset of Fig. 3(a), the
time dependence of npp for (Gly)20 is non-monotonous whereas the npw data in the inset of Fig. 3(b) exhibit a jump
at early time before reaching saturation. This again confirms the hydration effects for smaller chains. The overlap of
the hydrogen-bond kinetics for large N indicates that the collapse is not guided by the intra-peptide hydrogen bonds
but depends mostly on the intra-peptide van der Waals interactions.
However, the overlap of the hydrogen-bond data does not allow one to calculate the collapse time τc from the
time evolution of this quantity. More suitable for this purpose is the decay of the average squared radius of gyration
0.11101000.01100 0 ns 2 ns 5 ns 20 ns 05101520251030run 1run 2run 3run 4run 5S(q)t [ns]Rg2(t) [nm2](b)q [nm-1]q-4q-5/3(c)356
FIG. 3. Kinetics of H-bonding and scaling of the collapse time. (a) Time dependence of the number of protein-protein
hydrogen bonds npp(t) during collapse of (Gly)N for different N . To make the curves fall within the same scale the data is
normalized with npp(tf ); tf is the maximum run time the simulations are done. The inset shows time dependence of npp(t) for
(Gly)20. (b) Same as in (a) but for the number of protein-water hydrogen bonds npw(t). The inset again is same as the inset
of (a) but for npw(t). (c) Variation of the average squared radius of gyration R2
g(t) with time for the same systems that are
presented in (a) and (b). The solid black lines are respective fits using the form (1) and the corresponding β obtained is shown
as a function of N in the inset. (d) Dependence of the collapse times τc, extracted from the time decay of R2
of residue N . The solid line represents the behavior τc ∼ N z with z = 0.5.
g on the number
R2
g depicted in Fig. 3(c). The non-overlapping data are consistent with the respective solid lines obtained from the
previously proposed fit [16, 17]
g(t) = b0 + b1 exp[−(t/τc)β],
R2
(1)
where b0 corresponds to the value of R2
g(t) in the collapsed state, and b1 and β are associated non-trivial fitting
parameters. The obtained values of β [see the inset of Fig. 3(c)] indicate a very weak dependence on N , similar to
the case of the earlier studied collapse of synthetic homopolymers [16]. Although the above fit yields a collapse time
τc, more accurate estimates can be calculated from the time when R2
g(t) has decayed to 50% of its total decay, i.e.,
7
g(0) − R2
g = R2
∆R2
g(tf ). We plot the measured values of τc for different chain length N (including N = 20) in Fig.
3(d) to check for a scaling of the form τc ∼ N z. Due to the competition between hydration effects and intra-peptide
interactions that dominate for smaller N one expects distinct scaling forms for small and large N . Our data indeed
hint at the existence of two such scaling regions. Especially interesting is the consistency of our data for large N
with the solid line having z = 0.5. This exponent suggests that the dynamics is faster than the one observed in
MC simulations of non-biological homopolymer [16]. Surprisingly, it is even faster than in the case of homopolymer
collapse in presence of hydrodynamics [22, 23]. We conjecture that the more rapid collapse is due to the almost
instantaneous presence of intra-chain hydrogen bonds that hasten local ordering. Simulations of longer chains would
be desirable to confirm the value of z = 0.5 and the super-fast collapse mechanism in hydrogen-bonded polymers,
however, such simulations were computationally too costly to be considered in the present study.
In a final step we want to quantify the coarsening kinetics of the "pearl-necklace" observed in Fig. 2(a). A measure
of the relevant length scale, i.e., the mean cluster or pearl size Cs(t), can be obtained from a box-plot analysis of
the contact maps [27]. Conjecturing that the collapse is driven by the intra-peptide van der Waals attraction of the
backbone, we extract Cs(t) from an analysis of the contact probability P (cij) as a function of the contour distance
cij = i − j between any two Cα-atoms at the i-th and j-th position along the chain [33]. Two Cα-atoms are said
to have a contact if they are within a cut-off distance rc. Using rc = 2.5 nm, we show in Fig. 4(a) values of P (cij)
calculated at different times during the collapse of (Gly)100. These contact probabilities indicate indeed a growing
length scale as their decay slows with time. At the beginning, for t = 0 ns, the chain is in extended state and P (cij)
decays according to a power law P (cij) ∼ c
−γ
ij with an exponent γ = 1.5, as expected in a good solvent [34]. As time
progresses, this power-law behavior appears at larger cij after crossing over from a plateau-like behavior for small
cij which marks the local ordering along the chain. For any reasonable choice of rc the form of the curves stays
unchanged as demonstrated in Fig. 4(b). Similarly, the form of the curve also does not depend on the chain length
N as illustrated in the inset of Fig. 4(b) where we use rc = 2.5 nm and choose the point in time t = 2 ns.
The crossover point in the decay P (cij) as a function of cij is estimated from the discrete local slope as calculated
by [33]
γt(cij) =
∆ ln[P (cij)]
∆ ln[cij]
.
(2)
8
FIG. 4. Cluster growth during the collapse. (a) Contact probability P (cij) calculated using the cut-off rc = 2.5 nm, as
a function of the distance cij along the chain, at five different times during collapse of (Gly)100. (b) Shows the consistency or
the proportionality behavior of estimated contact probability P (cij) at a fixed time t = 2 ns using different rc as indicated.
The inset shows P (cij) at t = 2 ns using rc = 2.5 nm for different N . (c) The discrete slope γt obtained from Eq. (2) as a
function of cij for the times presented in (a). The solid line is for γt = 1. (d) The main frame shows the growth of the mean
cluster or pearl size Cs(t) with time for different N . The solid lines represent power-law behavior Cs(t) ∼ tα with α = 1 and
s (tp) as function of the shifted time tp = t − t0 for two different choices of t0.
2/3, respectively. The inset shows the plot of C p
The solid line there represents a linear behavior.
Plots of γt(cij) as a function of cij are shown Fig. 4(c) for the data presented in Fig. 4(a). The crossing of the data
with the γt = 1 line happens at larger cij as t increases, and thus this crossover point gives a measure of Cs(t). The
obtained Cs(t) for three different N are shown as a function of t on a double-log scale in the main frame of Fig. 4(d).
The flattening of the data at very large t is due to finite-size effects when no more ordering is possible due to formation
9
of single globule. At large t, before hitting size effects, the growth resembles a power law Cs(t) = AN tα where the
amplitude AN depends still on the chain length N as the considered N are not large enough. Hence, P (cij) calculated
using the same rc will overlap with each other, a fact that is demonstrated in the inset of Fig. 4(b). However, since
their form stays invariant in the large t regime, they apparently follow the same power law. Our data do not span
over decades in time, hence, it is hard to distinguish between α = 2/3 and 1 behavior as shown by the dashed and the
solid lines, respectively. In such cases it is advantageous to describe the growth instead as Cs(t) = C(t0) + AN (t− t0)α
by considering a crossover time t0 and cluster size C(t0). This ansatz, originally developed for ferromagnets [35], was
already necessary in our earlier work for describing the collapse of synthetic homopolymers [14, 16, 17]. Using the
p , with the shifted time tp = t − t0. For α = 1, the
s (tp) = Cs(t) − C(t0) one finds C p
transformation C p
s (tp) = Atα
above transformation is invariant under any choice of t0 in the post-crossover regime. This is demonstrated in the
inset of Fig. 4(d). The consistency of our data with the solid line representing a linear behavior further consolidates
our finding of a linear cluster growth.
In summary, we have investigated the nonequilibrium pathways by which polyglycine collapses in water. For short
chains, the pathway has few noticeable features and is driven by the competition between the hydration of the
peptide, opposing the collapse, and the intra-peptide attractions, favoring the collapse [7]. For long enough chains the
importance of hydration effects decreases, and the kinetics of hydrogen bonds indicates that van der Waals interactions
of the backbone dominate and drive the collapse. The nonequilibrium intermediates seen during the collapse exhibit
local ordering or clustering that is analogous to the phenomenological "pearl-necklace" picture known to be valid for
the earlier studied coarse-grained homopolymer models [20]. Using the contact probability of the Cα-atoms in the
backbone, we extract a relevant dynamic length scale, i.e., cluster size, that as in simple homopolymer models grows
linearly with time [16]. We believe that this linear growth is a result of the Brownian motion of the clusters and
subsequent coalescence as in the case of droplet growth in fluids [36].
Especially intriguing is that the scaling of the collapse time with length of the chain indicates a faster dynamics,
with a critical exponent z = 0.5 instead of z = 1, as seen in earlier homopolymer collapse studies [22, 23] that
considered simplified models describing non-hydrogen-bonded polymers such as polyethylene and polystyrene [37].
The smaller exponent suggests that the fast folding times of proteins (typically in the µs -- ms range for proteins
with ≈ 100 − 200 residues) may be connected with a mechanism that allows in amino acid based polymers a more
rapid collapse than seen in non-biological homopolymers, where collapse times of ≈ 300 ms [38] up to ≈ 350s [39] for
poly(N-isoporpylacrylamide) and polystyrene, respectively, have been reported. This would also have implications for
10
possible folding mechanisms as the fast collapse times in our simulations are connected with a quick appearance of
local ordering. In fact, we conjecture that the smaller exponent z is characteristic for collapse transitions where the
presence of intra-chain hydrogen bonding in amino acid based polymers immediately seeds (transient) local ordering,
a step that in non-hydrogen-bonded polymers only happens as the consequence of diffusive motion. However, in order
to test this conjecture, one would need to repeat first our above investigation for the other 19 amino acids. While such
study is beyond the scope of our current paper, the presented results demonstrate already that our approach provides
a general platform to understand various conformational transitions that occur in biomolecules via local ordering.
Another example would be, for instance, the helix-coil transition of polyalanine where the short-time dynamics has
already been explored [40, 41], or the study of two-time properties such as aging and dynamical scaling in collapse
and folding [15, 17].
METHODS
We construct (Gly)N molecules with hydrogenated N-terminus ( -- NH2) and C-terminus ( -- COOH). All-atom MD
simulations are performed using standard GROMACS 5.0.2 tools while CHARMM22 with CMAP corrections [42, 43]
is used for interactions between the atoms. For studying the collapse dynamics, we first prepare an extended chain
in gas phase at 1500 K. This follows solvation of this extended chain in a simple cubic box with water (modeled by
the TIP3P model [44]). The final MD run is performed at the desired quench temperature Tq = 290 K which is lower
than 310 K, roughly the collapse transition temperature of (Gly)N in water. The size of the box and the number of
water molecules, of course, are dependent on N and are so chosen that the number density of water molecules is same
for all N . For the smallest N , i.e., for N = 20 the default box size was 4.2 nm. Subsequently, the box sizes for longer
chains were determined using the relation Rg ∼ N 3/5. The size of the boxes should not have much role in collapse
provided the two ends of the chain do not interact while using the periodic boundary condition. However, the number
density of water molecules is supposed to play a role which we kept the same for all N . For N = 20 the total number
of water molecules used was ≈ 2000 giving a number density of ≈ 32 per nm−3 which was maintained for all N . After
the solvation we run our MD simulations using the Verlet-velocity integration scheme with time step δt = 2 fs, in the
NVT ensemble using the Nos´e-Hoover thermostat that conserves the linear momentum, and thus is believed to be
sufficient for preserving hydrodynamic effects [45]. Here, we use chains of length N ∈ [20, 50, 75, 100, 150, 200], and
all the results presented are averaged over 50 different initial configurations, except for N = 200 where this number is
15. The simulations are run up to time tf which is 10 ns for N = 20, 20 ns for N ∈ [50, 150], and 25 ns for N = 200.
For a polymer of length N (number of monomers) the squared radius of gyration is calculated as R2
11
g = Σij((cid:126)ri −
(cid:126)rj)2/2N 2. For (Gly)N the chain length was determined from N , the number of residues or repeating units which
contain a fixed set of atoms. Thus Rg for (Gly)N was calculated considering all the atoms present in all the residues.
However, the scaling can still be checked in terms N , as is done here. The structure factor for a polymer of length N
is calculated using the relation S((cid:126)q) = (1/N )Σij exp(−i(cid:126)q · (cid:126)rij), where (cid:126)rij is the distance vector between the i-th and
j-th monomer along the chain. As explained above in the case for measuring Rg, for S(q) too we use all the atoms in
all the residues. We calculate the hydrogen bonds using the standard GROMACS tool gmx hbond. It considers all
possible donars and acceptors and decides for the existence of a hydrogen bond if the distance between them is less
than 0.35 nm and the hydrogen-donar-acceptor angle is less than 30°.
ACKNOWLEDGMENTS
This project was funded by the Deutsche Forschungsgemeinschaft (DFG) under Grant Nos. SFB/TRR 102 (project
B04) and JA 483/33-1 and the National Institutes of Health (NIH) under grants GM120578 and GM120634. It was
further supported by the Deutsch-Franzosische Hochschule (DFH-UFA) through the Doctoral College "L4" under
Grant No. CDFA-02-07 and the Leipzig Graduate School of Natural Sciences "BuildMoNa". U.H. thanks the Institut
fur Theoretische Physik and especially the Janke group for kind hospitality during his sabbatical stay at Universitat
Leipzig.
CONTRIBUTIONS
All authors contributed to develop the project and S.M. performed the simulations. All authors discussed, analyzed
the results and wrote the manuscript.
COMPETING FINANCIAL INTERESTS
The authors declare that they have no competing financial interests.
[1] W.H. Stockmayer, "Problems of the statistical thermodynamics of dilute polymer solutions," Macromol. Chem. Phys. 35,
54 (1960).
12
[2] I. Nishio, S.-T. Sun, G. Swislow, and T. Tanaka, "First observation of the coil -- globule transition in a single polymer
chain," Nature 281, 208 (1979).
[3] L. Pollack, M.W. Tate, A.C. Finnefrock, C. Kalidas, S. Trotter, N.C. Darnton, L. Lurio, R.H. Austin, C.A. Batt, S.M.
Gruner, and S.G.J. Mochrie, "Time resolved collapse of a folding protein observed with small angle x-ray scattering,"
Phys. Rev. Lett. 86, 4962 (2001).
[4] M. Sadqi, L.J. Lapidus, and V. Munoz, "How fast is protein hydrophobic collapse?" Proc. Natl. Acad. Sci. U.S.A. 100,
12117 (2003).
[5] C.J. Camacho and D. Thirumalai, "Kinetics and thermodynamics of folding in model proteins," Proc. Natl. Acad. Sci.
U.S.A. 90, 6369 (1993).
[6] G. Reddy and D. Thirumalai, "Collapse precedes folding in denaturant-dependent assembly of ubiquitin," J. Phys. Chem.
B 121, 995 (2017).
[7] D. Asthagiri, D. Karandur, D.S. Tomar, and B.M. Pettitt, "Intramolecular interactions overcome hydration to drive the
collapse transition of gly15," J. Phys. Chem. B 121, 8078 (2017).
[8] W. Kauzmann, "Some factors in the interpretation of protein denaturation1," Adv. Protein Chem. 14, 1 (1959).
[9] K.A. Dill, "Dominant forces in protein folding," Biochemistry 29, 7133 (1990).
[10] D.W. Bolen and G.D. Rose, "Structure and energetics of the hydrogen-bonded backbone in protein folding," Ann. Rev.
Biochem. 77, 339 (2008).
[11] H.T. Tran, A. Mao, and R.V. Pappu, "Role of backbone-solvent interactions in determining conformational equilibria of
intrinsically disordered proteins," J. Am. Chem. Soc. 130, 7380 -- 7392 (2008).
[12] L.M.F. Holthauzen, J. Rosgen, and D.W. Bolen, "Hydrogen bonding progressively strengthens upon transfer of the protein
urea-denatured state to water and protecting osmolytes," Biochemistry 49, 1310 (2010).
[13] D.P. Teufel, C.M. Johnson, J.K. Lum, and H. Neuweiler, "Backbone-driven collapse in unfolded protein chains," J. Mol.
Bio. 409, 250 (2011).
[14] S. Majumder and W. Janke, "Cluster coarsening during polymer collapse: Finite-size scaling analysis," Europhys. Lett.
110, 58001 (2015).
[15] S. Majumder and W. Janke, "Evidence of aging and dynamic scaling in the collapse of a polymer," Phys. Rev. E 93,
032506 (2016).
[16] S. Majumder, J. Zierenberg, and W. Janke, "Kinetics of polymer collapse: Effect of temperature on cluster growth and
aging," Soft Matter 13, 1276 (2017).
[17] H. Christiansen, S. Majumder, and W. Janke, "Coarsening and aging of lattice polymers: Influence of bond fluctuations,"
J. Chem. Phys. 147, 094902 (2017).
[18] A.J. Bray, "Theory of phase-ordering kinetics," Adv. Phys. 51, 481 (2002).
13
[19] P.-G. de Gennes, "Kinetics of collapse for a flexible coil," J. Phys. Lett. 46, 639 (1985).
[20] A. Halperin and P.M. Goldbart, "Early stages of homopolymer collapse," Phys. Rev. E 61, 565 (2000).
[21] A. Byrne, P. Kiernan, D. Green, and K.A. Dawson, "Kinetics of homopolymer collapse," J. Chem. Phys. 102, 573 (1995).
[22] C.F. Abrams, N.K. Lee, and S.P. Obukhov, "Collapse dynamics of a polymer chain: Theory and simulation," Europhys.
Lett. 59, 391 (2002).
[23] N. Kikuchi, J.F. Ryder, C.M. Pooley, and J.M. Yeomans, "Kinetics of the polymer collapse transition: The role of
hydrodynamics," Phys. Rev. E 71, 061804 (2005).
[24] G. Reddy and A. Yethiraj, "Implicit and explicit solvent models for the simulation of dilute polymer solutions," Macro-
molecules 39, 8536 (2006).
[25] J. Guo, H. Liang, and Z.-G. Wang, "Coil-to-globule transition by dissipative particle dynamics simulation," J. Chem.
Phys. 134, 244904 (2011).
[26] A. Montesi, M. Pasquali, and F.C. MacKintosh, "Collapse of a semiflexible polymer in poor solvent," Phys. Rev. E 69,
021916 (2004).
[27] A. Lappala and E.M. Terentjev, "Raindrop coalescence of polymer chains during coil -- globule transition," Macromolecules
46, 1239 (2013).
[28] D.K. Wilkins, S.B. Grimshaw, V. Receveur, C.M. Dobson, J.A. Jones, and L.J. Smith, "Hydrodynamic radii of native
and denatured proteins measured by pulse field gradient nmr techniques," Biochemistry 38, 16424 (1999).
[29] V.N. Uversky, "Natively unfolded proteins: A point where biology waits for physics," Protein Sci. 11, 739 (2002).
[30] I.R. Cooke and D.R.M. Williams, "Collapse dynamics of block copolymers in selective solvents: Micelle formation and the
effect of chain sequence," Macromolecules 36, 2149 (2003).
[31] T.T. Pham, B. Dunweg, and J.R. Prakash, "Collapse dynamics of copolymers in a poor solvent: influence of hydrodynamic
interactions and chain sequence," Macromolecules 43, 10084. (2010).
[32] M. Rubenstein and R.H. Colby, Polymer Physics (Chemistry) (Oxford University Press, Oxford, 2003).
[33] V.F. Scolari, G. Mercy, R. Koszul, A. Lesne, and J. Mozziconacci, "Kinetic signature of cooperativity in the irreversible
collapse of a polymer," Phys. Rev. Lett. 121, 057801 (2018).
[34] P.-G. de Gennes, Scaling Concepts in Polymer Physics (AIP, Melville, New York, 1980).
[35] S. Majumder and S.K. Das, "Domain coarsening in two dimensions: Conserved dynamics and finite-size scaling," Phys.
Rev. E 81, 050102 (2010).
[36] K. Binder and D. Stauffer, "Theory for the slowing down of the relaxation and spinodal decomposition of binary mixtures,"
Phys. Rev. Lett. 33, 1006 (1974).
[37] K. Kremer and G.S. Grest, "Dynamics of entangled linear polymer melts: A molecular-dynamics simulation," J. Chem.
Phys. 92, 5057 (1990).
14
[38] J. Xu, Z. Zhu, S. Luo, C. Wu, and S. Liu, "First observation of two-stage collapsing kinetics of a single synthetic polymer
chain," Phys. Rev. Lett. 96, 027802 (2006).
[39] B. Chu, Q. Ying, and A.Yu. Grosberg, "Two-stage kinetics of single-chain collapse. Polystyrene in cyclohexane," Macro-
molecules 28, 180 (1995).
[40] E. Arashiro, J.R. Drugowich de Fel´ıcio, and U.H.E. Hansmann, "Short-time dynamics of the helix-coil transition in
polypeptides," Phys. Rev. E 73, 040902 (2006).
[41] E. Arashiro, J.R. Drugowich de Fel´ıcio, and U.H.E. Hansmann, "Short-time dynamics of polypeptides," J. Chem. Phys.
126, 045107 (2007).
[42] A.D. MacKerell Jr., D. Bashford, M.L.D.R. Bellott, R.L. Dunbrack Jr., J.D. Evanseck, M.J. Field, S. Fischer, J. Gao,
H. Guo, S. Ha, et al., "All-atom empirical potential for molecular modeling and dynamics studies of proteins," J. Phys.
Chem. B 102, 3586 (1998).
[43] A.D. Mackerell Jr., M. Feig, and C.L. Brooks III, "Extending the treatment of backbone energetics in protein force fields:
limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics
simulations," J. Comp. Chem. 25, 1400 (2004).
[44] W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, R.W. Impey, and M.L. Klein, "Comparison of simple potential functions
for simulating liquid water," J. Chem. Phys. 79, 926 (1983).
[45] D. Frenkel and B. Smit, Understanding Molecular Simulations: From Algorithms to Applications (Academic Press, San
Diego, 2002).
|
1606.07978 | 2 | 1606 | 2017-02-16T15:20:24 | String networks with junctions in competition models | [
"physics.bio-ph"
] | In this work we give specific examples of competition models, with six and eight species, whose three-dimensional dynamics naturally leads to the formation of string networks with junctions, associated with regions that have a high concentration of enemy species. We study the two- and three-dimensional evolution of such networks, both using stochastic network and mean field theory simulations. If the predation, reproduction and mobility probabilities do not vary in space and time, we find that the networks attain scaling regimes with a characteristic length roughly proportional to $t^{1/2}$, where $t$ is the physical time, thus showing that the presence of junctions, on its own, does not have a significant impact on their scaling properties. | physics.bio-ph | physics | Physics Letters A 00 (2018) 1–10
Physics
Letters A
String networks with junctions in competition models
P. P. Avelinoa,b, D. Bazeiac, L. Losanoc, J. Menezese,d, B. F. de Oliveiraf
bDepartamento de F´ısica e Astronomia, Faculdade de Ciencias, Universidade do Porto, Rua do Campo Alegre 687, 4169-007 Porto, Portugal
aCentro de Astrof´ısica da Universidade do Porto, Rua das Estrelas, 4150-762 Porto, Portugal
cDepartamento de F´ısica, Universidade Federal da Para´ıba, 58051-970 Joao Pessoa, PB, Brazil
dEscola de Ciencias e Tecnologia, Universidade Federal do Rio Grande do Norte
Caixa Postal 1524, 59072-970 Natal, RN, Brazil
eInstitute for Biodiversity and Ecosystem Dynamics, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
fDepartamento de F´ısica, Universidade Estadual de Maring´a, Av. Colombo, 5790, 87020-900 Maring´a, PR, Brazil
Abstract
In this work we give specific examples of competition models, with six and eight species, whose three-dimensional dynamics
naturally leads to the formation of string networks with junctions, associated with regions that have a high concentration of enemy
species. We study the two- and three-dimensional evolution of such networks, both using stochastic network and mean field theory
simulations. If the predation, reproduction and mobility probabilities do not vary in space and time, we find that the networks
attain scaling regimes with a characteristic length roughly proportional to t1/2, where t is the physical time, thus showing that the
presence of junctions, on its own, does not have a significant impact on their scaling properties.
c(cid:13) 2011 Published by Elsevier Ltd.
Keywords: population dynamics, string networks
1. Introduction
Competition models are widely regarded as a crucial tool to understand the mechanisms leading to biodiversity [1–
4] (see also [5, 6] for a review). Although the simplest competition models usually consider three species and allow for
an equal number of basic microscopic actions (motion, reproduction and predation), many interesting generalizations,
including further species and more complex interaction rules, have been considered in the literature [7–31]
In [14, 15], a broad family of spatial stochastic May-Leonard models with an arbitrary number of species has
been introduced. There, many interesting features of this family of models, including the dynamics of complex
networks of spiralling patterns and interfaces, have been investigated in detail. Recently, in [25], it has been shown that
specific sub-classes of this family of models may lead to the emergence of string networks in three spatial dimensions.
These strings are associated to predator-prey interactions which occur mainly along lines corresponding to a high
concentration of enemy species. The strings studied in [25] do not have junctions and their dynamics has been shown
to be curvature driven.
In this paper we consider the emergence of string networks with junctions in the context of specific spatial stochas-
tic competition models belonging to the general family introduced in [14, 15]. We investigate the two- and three-
dimensional dynamics of these models using both stochastic and mean field network simulations. The outline of this
paper is as follows. In Set. 2 we introduce a sub-class of stochastic May-Leonard models allowing for the formation
of string networks with junctions in three spatial dimensions. In Sec. 3 we investigate the stochastic evolution of two-
1
7
1
0
2
b
e
F
6
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
8
7
9
7
0
.
6
0
6
1
:
v
i
X
r
a
/ Physics Letters A 00 (2018) 1–10
2
and three-dimensional networks in models belonging to the above sub-class with N = 6 or N = 8 species. In Sec. 4
we study the dynamics of such systems using mean field theory simulations, considering also the particular case of
the collapse of a circular loop. In Sec. 5 we constrain the scaling parameter λ governing the macroscopic dynamics
of these models, considering various choices for the mobility, predation and reproduction parameters. Finally, we
conclude in Sec. 6.
2. Family of models
Here, we consider a sub-class of the more general family of spatial stochastic May-Leonard models introduced in
Refs. [14, 15]. We investigate models with an even number of species N > 4, where each species competes with N − 4
other species. The competition diagrams for the simplest cases with N = 6 and N = 8 species are illustrated in the
left and right panels of Fig. 1, respectively. The double arrows indicate that the predation between competing species
is bi-directional. Except for the labelling of the different species, the diagrams depicted in Fig. 1 are invariant under
rotations by θn = 2 π(n/N), with n = 0,±1,±2, . . . ,±(N − 1), indicating the presence of a ZN symmetry.
In these models individuals of N species are initially distributed on square or cubic lattices with N sites. The
different species are labeled by i = 1, . . . , N, and the cyclic identification i = i + k N, where k is an integer, is made.
The sum of the number of individuals of the species i (Ii) with the number of empty sites (IE) is equal to the total
number of sites (N), that is
N(cid:88)
Ii + IE = N.
(1)
i=1
At each time step a random individual (active) is chosen to interact with one of its nearest neighbours (passive),
also selected randomly. The number of neighbours is equal to four, in the case of the two-dimensional square lattice,
and to six, in the case of the three-dimensional cubic lattice. The unit of time ∆t = 1 is defined as the time necessary for
N interactions to occur (one generation). The possible interactions are classified as Motion i (cid:12) → (cid:12) i , Reproduction
i ⊗ → ii , or Predation i (i±α) → i ⊗ , where (cid:12) may be any species (i) or an empty site (⊗) and α = 1, . . . , (N−4)/2. A
constant predation probability p between competing species, and constant reproduction and mobility probabilities (r
and m, respectively), common to all species, is considered. Although this class of models is defined for a generic even
N > 4, in this paper we shall investigate explicitly the models illustrated in Fig. 1, with N = 6 and N = 8 different
species.
3. Stochastic network simulations
We performed a series of stochastic network simulations of the models with six and eight species in two and
three spatial dimensions (in square and cubic lattices, respectively). The numerical results show that, as soon as
the simulations start, individuals of the same species, originally distributed randomly throughout the lattice, share
common spatial regions. The individuals tend not to be close to competitors, but in the vicinity of individuals of the
same species or of other neutral species.
In two spatial dimensions such spatial configurations promote the coexistence of species which may be organized
either clockwise or counterclockwise, around roughly circular regions with a significantly higher density of empty
sites. Attacks and counter-attacks between competing species on opposite sides of the battle cores ensure the stability
of such spatial patterns. Similar two-dimensional arrangements of the species have been found in Ref. [25]. These can
be described as defect/anti-defect configurations associated to clockwise/counterclockwise vortex states, respectively.
The average area of the defect and anti-defect cores depends on the interaction probabilities (m, p, r), but it is roughly
constant in time.
In contrast with the model described in Ref. [25], where every species would compete with N−3 distinct ones, here
each species has less competitors (each species has N − 4 competitors and three neutral species). As a consequence,
the number of configurations which promote coexistence among species is enlarged.
In fact, while in Ref. [25]
all N species have to be gathered around the defect cores in order to guarantee their stability, here, a stable defect
configuration requires only four species. Defects/anti-defects arise when the different species dispose themselves in
the clockwise/counterclockwise vortex configurations shown in Fig. 2 {i, i + N
2 (note
2 }, where i = 1, ..., N
2 , i + 1, i − N−2
2
/ Physics Letters A 00 (2018) 1–10
3
Figure 1. Diagrams describing the possible predator-prey interactions in the models with N = 6 (left panel) and N = 8 (right panel) different
species.
Figure 2. Illustration of a defect/anti-defect pair with four species in a model with an arbitrary number of species N, where i = 1, ..., N
2 .
that each species i does not interact with the species i ± N−2
kinds of defect/anti-defects with four species.
2 and i + N
2 ). These configurations account for N
2 different
Consider the competing species i and i + 1 belonging to the defect/anti-defect configuration shown in Fig. 2. The
average number of attacks per unit time from individuals of the outer species i + 1 supersedes those from individuals
of the inner species i. This implies that individuals of the outer species tend to invade the territory of the inner ones
causing an approximation and annihilation of the defect/anti-defect pair. We shall show that the defect/anti-defect
cores attract each other, having a velocity which is, on average, inversely proportional to the distance between them.
In contrast, a pair of clockwise (or counterclockwise) defects (anti-defects) cannot annihilate and repel each other.
Moreover, defect/anti-defect configurations with a larger number of species may arise. In fact, defect/anti-defect
species, in which each species competes with n − 3 other species, may appear in
configurations with n = 4, ..., N+2
2
models with an even number N > 6 of species.
All network simulation snapshots presented throughout this paper were obtained by assuming m = 0.10, p = 0.80
and r = 0.10. Although these values were found to be adequate for visualization purposes, we verified that many other
choices of the parameters would provide similar qualitative results.
3.1. 6 species
Let us focus on the simplest case with N = 6. The left panel of Fig. 3 shows one snapshot taken from a
two-dimensional 10242 stochastic network simulation with periodic boundary conditions. Each grid point is at most
occupied by one individual which belongs to the species indicated by the same color as in Fig. 1. In addition, empty
spaces are represented by white dots.
The spatial patterns show that all defects involve four different species. Most of the predator-prey interactions take
place at the defect core, which is surrounded by two pairs of competing domains, each domain being dominated by a
single species. More specifically, one may identify three types of defects-anti/defects where domains dominated by
the species {i, i − 1, i + 1, i − 2} (where i = 1, 2, 3) arrange themselves in this order (or the reverse) around the defect
(anti-defect) cores. Due to the periodic boundary conditions, the number of defects in each simulation is equal to
the number of anti-defects. As the network evolves, the number of defects becomes increasingly smaller due to the
annihilation of defect/anti-defect pairs. Ultimately, at large enough t, only one group of three species which do not
3
12345612345678/ Physics Letters A 00 (2018) 1–10
4
Figure 3. Snapshots obtained from two-dimensional 10242 stochastic network simulations of the N = 6 (left panel) and N = 8 (right panel) models.
The color scheme used here is the same as represented in Fig. 1 for i = 1.
Figure 4. Snapshots taken from a three-dimensional 643 region of a 2563 stochastic network simulation of the N = 6 (left panel) and N = 8 (right
panel) models.
compete each other may survive (either (1, 3, 5) or (2, 4, 6)), and the symmetry of the model is said to be spontaneously
broken.
We also run a series of 2563 three-dimensional stochastic network simulations with periodic boundary conditions
of the N = 6 model. We notice that the extension to three spatial dimensions of the dynamics presented in the left
panel of Fig. 3 gives rises to a string network with Y-type junctions. Such strings represent regions with a significant
larger number density of empty spaces, which appear as a consequence of the frequent predation interaction between
competing species taking place at their core.
The snapshot shown in the left panel of Fig. 4 represents a 643 region of the entire 2563 three-dimensional lattice.
It presents the contour plots associated to a fixed value of the density of empty sites, which highlight the presence of
a string network with junctions (note that, in order to improve the visualization, the number density of empty spaces
has been convolved with a gaussian filter function).
Throughout their evolution the strings intersect and intercommute (exchange partners). This process is responsible
for the production of string loops which collapse with a characteristic velocity roughly proportional to the loop charac-
teristic scale. The collapse of a string loop is curvature driven and it is associated to the existence of a defect/anti-defect
pair on any plane intersecting the loop. The loop collapse will be further considered in Sec. 4.3.
Since any planar defect or anti-defect involves only four species, each species joins two different strings in three
spatial dimensions. This is responsible for the appearance of Y-type junctions, defined as the meeting point of three
different strings. The stability of the junctions is ensured by predator-prey interactions, given that each species has
competitors in two different strings.
The upper panel of Fig. 5 illustrates the different possible configurations of a string junction in the N = 6 model,
where i = 1, 2, 3. Note that, in this model, all the species are involved in the formation of a Y-type junctions, each one
of them being associated to two different strings.
3.2. 8 species
The results obtained for N = 8 are qualitatively similar to those presented for N = 6, but with an increased
In this case, apart from the four different defects/anti-
complexity associated with the larger number of species.
4
/ Physics Letters A 00 (2018) 1–10
5
Figure 5. Illustration of the formation of Y-type junctions in the models N = 6 (left panel) and N = 8 (right panel). The cross-sections of the
different types of strings joining the junctions are presented, where i = 1, ..., N
2 .
defects with four species represented by {i, i + 4, i + 1, i − 3} (where i = 1, ..4), there are eight defect/anti-defect
configurations composed by five species. They are composed by the clockwise/counterclockwise disposition of the
species {i, i + 4, i + 1, i − 2, i + 3} (where i = 1, ..8) around the defect cores. We call the defects formed by four and
five species, type I and II, respectively.
The right panel of Fig. 3 depicts one snapshot taken from a 10242 stochastic network simulation of the N = 8
model, with periodic boundary conditions. The colors indicate the species that individuals belong to (See Fig. 1).
White dots represent empty sites.
In addition, the results provided by three-dimensional simulations are presented in the right panel of Fig. 4. The
snapshot represents a 643 region of a three-dimensional 2563 stochastic network simulations with periodic boundary
conditions of the N = 8 model.
Analogously to the N = 6 model, the extension to three spatial dimensions gives rises to a string network with
junctions. Nonetheless, in the N = 8 model there are two different types of strings, in which either four (type I) or
five different species (type II) are involved. One string of type I and two of type II are required in order to produce the
stable Y-type junctions seen in the simulations of the N = 8 model. The lower panel of Fig. 5 illustrates the formation
of a Y-type junction in this model.
4. Mean field theory simulations
Let us define N + 1 scalar fields (φ0, φ1, φ2, . . ., φN) representing the fraction of space around a given point
occupied by empty sites (φ0) and by individuals of the species i (φi), satisfying the constraint φ0 + φ1 + . . . + φN = 1.
For an even N > 4 the mean field equations of motion
(2)
(3)
φ0 = D∇2φ0 − rφ0
i=1
φi = D∇2φi + rφ0φi − p
N(cid:88)
φi + p
N−4
2(cid:88)
α=1
N−4
2(cid:88)
α=1
N(cid:88)
i=1
N−1(cid:88)
α= N+4
2
φiφi+α +
φiφi+α
φiφi+α +
φiφi+α
,
N−1(cid:88)
,
α= N+4
2
describe the average dynamics of the models studied in the previous section (see [7], for more details). Here, a dot
represents a derivative with respect to the physical time and D = 2m is the diffusion rate.
4.1. Two and three-dimensional numerical results
Figure 6 depicts two snapshots taken from 10242 mean field network simulations with periodic boundary condi-
tions of the N = 6 (left panel) and N = 8 (right panel) models. Initial conditions with φi = 1 if i = s and φi = 0 if
5
/ Physics Letters A 00 (2018) 1–10
6
Figure 6. Snapshots obtained from two-dimensional 10242 mean field simulations of the N = 6 (left panels) and N = 8 (right panels) models.
Figure 7. Snapshots taken from three-dimensional mean field simulations of the N = 6 (left panel) and N = 8 (right panel) models. The figures
represent 643 regions of 2563 simulations.
i (cid:44) s were set at each grid point (φ0 was set to zero at every grid point). Here s is a species drawn randomly at each
grid point.
The results provided by the mean field simulations (Fig. 6) are similar to those obtained from the stochastic
network evolution (Fig. 3), except for the noise (the color scheme is the same in both figures). This correspondence
also occurs in the three-dimensional simulations. In particular, the macroscopic dynamics depicted in the snapshot
shown in Fig. 7 obtained from three-dimensional 2563 mean field theory simulations of the N = 6 and N = 8 models,
is similar (again, except for the noise) to that shown in Fig. 4.
4.2. Defect profile
Let us now consider the defect profiles, i.e., the stationary number density of empty spaces in and around the
defect cores. In other words, let us study the spatial distribution of φ0 around a defect center, considering that the
defect has radial symmetry and is centered at r = 0.
Furthermore, we aim to understand the dependence of the defect profile on the parameters (m, p, r). To this
purpose, we consider three different models characterized by a domination of motion [model M: (m, p, r)=(0.8, 0.1,
0.1)], predation [model P: (m, p, r)= (0.1, 0.8, 0.1)] or reproduction [model R: (m, p, r) = (0.1, 0.1, 0.8)] interactions.
Figure 8 shows the value of φ0, as a function of the distance r to the defect core, obtained for defects associated
with four and five species from mean field simulations of the N = 6 (upper panel) and N = 8 (middle and lower
panels) models. The disposition of the species around the defect cores is shown in the inset plots. The numerical
results show a strong dependence of the defect profile on the model parameters. Specifically, the defect profile height
is larger in the case of model P, dominated by predation interactions. On the contrary, a larger reproduction rate leads
to a smaller number density of empty spaces (model R). Finally, a larger mobility parameter leads to the broadening
of the defect profile, since the individuals are more likely to move outside the core of the defect into enemy territory
(model M).
Figure 8 also shows that, in the model with N = 8 species and for fixed (m, p, r), the defect is broader when the
number of species composing the defect is increased from 4 to 5.
6
/ Physics Letters A 00 (2018) 1–10
7
Figure 8. Defect profiles obtained for defects with four and five species from mean field simulations for N = 6 (left panel) and N = 8 (right and
lower panels). The inset plots represent the disposition of the species around the defect cores.
4.3. String loop
We now investigate the collapse of a circular string loop using mean field theory simulations in a cubic lattice.
To identify the string we define a new variable ϕ((cid:126)r, t) ≡ max(φ0((cid:126)r, t)) − φc
0 represents a threshold which
guarantees that only grid points with a high number density of empty sites, close to the core of the string, are identified
as belonging to the string. The average number density of empty sites associated to the string is then defined by
0, 0). Here, φc
(cid:88)
(cid:126)r
ρ(t) =
1
N 2
ϕ((cid:126)r, t) .
(4)
We recall that the average number of empty sites per unit string length (µ) does not change significantly with time.
Therefore, the loop perimeter is proportional to ρ(t) and the area of the string loop a(t) evolves proportionally to ρ2(t).
The time evolution of the area a(t) of the circle enclosed by the loop is determined using mean field theory
simulations in a 2563 cubic lattice. We run simulations for two different threshold values in order to ensure the
accuracy and reliability of the results. The upper panel of Fig. 9 displays the evolution of the area of a string loop
formed by four distinct species a4(t) in the N = 6 model. Note that the results are almost identical for both choices
of threshold values, φc
= 0.15, which correspond approximately to 25% and 60% of the maximum of
0
φ0 (φmax
) at the core of the string. Similar results were found for the evolution of area of the circle enclosed by string
loops associated to four and five distinct species a4(t) and a5(t) in the N = 8 model (see lower panel of Fig. 9). Fig.
9 shows that the loop area decreases linearly in time according to
1 − t
tc
= 0.06 and φc
0
a(t) = a0
(cid:33)
(5)
0
(cid:32)
7
,
00.050.10.150.20.250.3402002040φ0r6species00.0100.0700.300.3PMR1425PMR00.050.10.150.20.25402002040φ0r8species00.00800.0600.2500.25PMR1485PMR00.050.10.150.20.250.30.35402002040φ0r8species00.0100.08500.3500.35PMR14726PMR/ Physics Letters A 00 (2018) 1–10
8
Figure 9. Mean field evolution of the areas of string loops in the N = 6 (left panel) and N = 8 (right panel) models, computed by taking different
threshold values. Note that a4(t) and a5(t) represent the areas of type I and type II string loops, respectively.
Figure 10. The dependence of the scaling exponent λ on the threshold φc
species. The results were obtained by carrying out 10 simulations of 20482 two-dimensional networks for a wide range of φc
R. The error bars represent the standard deviation in an ensemble of 10 simulations.
0 in a mean field simulation with N = 6 (left panel) and N = 8 (right panel)
0, for models M, P and
where tc is the collapse time or, equivalently, that radius of curvature decreases proportionally to t1/2. A similar
result has been obtained in Ref. [15] for a different model allowing for string networks without junctions.
5. Scaling behavior
Finally, we consider the scaling behavior of the models investigated in the previous sections using mean field
theory simulations. To this purpose, let us define the characteristic length L of the defect network as
(cid:115)
(cid:114) µ
ρ
L =
∝
1
ρ
,
(6)
where µ is the average number of empty spaces associated with the defects (per unit string length, in three spatial
dimensions) and ρ is the average number density of empty sites associated to the defect network defined by Eq. (4).
By fitting a scaling law L ∝ tλ, we constrain the evolution of the characteristic length of the string network with
time for different sets of 10 mean field simulations of the models M, R and P. Each simulation starts with different
random initial conditions.
The upper and lower panels of Fig. 10 show the dependence of λ on the threshold φc
0 for models M, P and R, for
N = 6 and N = 8, respectively (in units of φc
is taken as the maximum of φ0 at the core of the
string of type I. The error bars represent the standard deviation in an ensemble of 10 simulations. Figure 10 shows
that if the φc
, the scaling constant λ does not show a significant dependence on the
threshold. Outside this interval, this is no longer the case. For lower values of φc
0, this happens because low density
0 is between 20% and 60% of φmax
0/φmax
0
). For N = 8, φmax
0
0
8
00.20.40.60.810600120018002400a4tφc0=0.06φc0=0.1500.20.40.60.811.2060012001800240000.20.40.60.81a4a5tφc0=0.07φc0=0.175φc0=0.07φc0=0.1750.350.400.450.500.550.600.6500.20.40.60.81λφc0/φmax0MRP6species-2D0.300.350.400.450.500.550.6000.20.40.60.81λφc0/φmax0MRP8species-2D/ Physics Letters A 00 (2018) 1–10
9
Figure 11. Scaling behavior for the models with N = 6 and N = 8 different species obtained using 20482 two-dimensional (left panel) and 2563
three-dimensional (right panel) mean field numerical simulations.
regions far away from the core are being taken as belonging to the defect. On the other hand, the number of lattice
points associated with the defect may become too small for higher values of φc
0.
The average evolution of L with time t was obtained for sets of 10 distinct two- and three-dimensional mean
field network simulations (20482 and 2563) with random initial conditions. Fig. 11 shows that the characteristic
lengths L6S (6 species) and L8S (8 species) evolve in reasonable agreement with the scaling law L ∝ tλ, with λ = 1/2,
characteristic of networks in which the dynamics is curvature driven. Here, the points denote the average value of
L computed from the simulation and the error bars provide information on the root-mean-square deviation for each
set of 10 simulations. In all cases the value of L was normalized to unity at t = 100, the parameters (m, p, r) were
set to (0.10, 0.80, 0.10) and the threshold φc
0 was obtained by
considering the value of φmax
0 was fixed at 40% of φmax
(for N = 8 the treshold φc
obtained for type II strings).
0
0
These results are consistent with those obtained in Ref. [25] for models allowing for string networks without
junctions in three spatial dimensions. A similar behavior may also be found in other physical systems, in particular in
the case of curvature driven dynamics of string networks in condensed matter.
6. Comments and Conclusions
In this work we have shown that there are specific sub-classes, with an even number of species, of a more general
family of May-Leonard models which lead to the formation of cosmic string networks with junctions, associated
to regions with a high concentration of empty spaces. We have investigated the dynamics of these networks using
stochastic and mean field network simulations, assuming that the predation, reproduction and mobility probabilities
are constant in space and time. We have found that the presence of junctions does not have a significant impact on
the scaling behaviour of the characteristic macroscopic scale of the network L with the physical time t, showing that
it grows roughly proportional to t1/2. We have also shown that our results are not strongly dependent on the specific
values of the mobility, predation or reproduction probabilities.
Our findings are expected to be relevant not only for the evolution of biological populations in three spatial
dimensions but also for the understanding of the evolution of cosmic string networks with junctions in other contexts.
In particular, in [32] one investigates bifurcation and pattern changing in the relativistic regime, suggesting how
to solve the cosmological domain wall problem. Also, in [33] it has been shown that, although the evolution of
the characteristic macroscopic length and velocity of interface networks with physical time in relativistic and non-
relativistic regimes is very different, single snapshots with the same characteristic scale do not clearly differentiate
between these two regimes. Hence, our results may also be relevant for the understanding of the dynamical behaviour
of complex relativistic defect networks with junctions in a cosmological setting (e.g. cosmic superstrings [34–36]).
Acknowledgements
We thank CAPES, CNPq, CNPq/Fapern, and FCT-Portugal for financial support. The work of PPA was supported
by Fundac¸ao para a Ciencia e a Tecnologia (FCT) through the Investigador FCT contract of reference IF/00863/2012
9
100101102103104100λp=0.40±0.01λp=0.41±0.01L6SL8St8S-φc0=0.146S-φc0=0.122D100101102103100λp=0.46±0.03λp=0.45±0.02L6SL8St8S-φc0=0.146S-φc0=0.123Dand POPH/FSE (EC) by FEDER funding through the program Programa Operacional de Factores de Competitivi-
dade, COMPETE.
/ Physics Letters A 00 (2018) 1–10
10
References
[1] B. Kerr, M. A. Riley, M. W. Feldman, B. J. M. Bohannan, Mobility promotes and jeopardizes biodiversity in rockpaperscissors games, Nature
[2] J. J. Leisner, J. Haaber, Intraguild predation provides a selection mechanism for bacterial antagonistic compounds, Proceedings of the Royal
[3] H. Cheng, N. Yao, Z. Huang, J. Park, Y. Do, Y. Lai, Mesoscopic interactions and species coexistence in evolutionary game dynamics of cyclic
418 (2002) 171.
Society B: Biological Sciences 279 (2012) 4513.
competitions, Scientific Reports 4 (2014) 7486.
[4] A. J. Daly, J. M. Baetens, B. D. Baets, The impact of initial evenness on biodiversity maintenance for a four-species in silico bacterial
community, Journal of Theoretical Biology 387 (2015) 189 – 205.
[5] R. C. Sole, J. Bascompte, Self-Organization in Complex Ecosystems, Princeton UP, Princeton, 2006.
[6] M. A. Nowak, Evolutionary Dynamics: Exploring the Equations of Life, Harvard University Press, 2006.
[7] R. May, W. Leonard, Nonlinear aspects of competition between three species, SIAM Journal on Applied Mathematics 29 (1975) 243.
[8] T. Reichenbach, M. Mobilia, E. Frey, Mobility promotes and jeopardizes biodiversity in rockpaperscissors games, Nature 448 (2007) 1046.
[9] T. Reichenbach, M. Mobilia, E. Frey, Noise and correlations in a spatial population model with cyclic competition, Phys. Rev. Lett. 99 (2007)
[10] M. Peltomaki, M. Alava, Three- and four-state rock-paper-scissors games with diffusion, Phys. Rev. E 78 (2008) 031906.
[11] W.-X. Wang, X. Ni, Y.-C. Lai, C. Grebogi, Pattern formation, synchronization, and outbreak of biodiversity in cyclically competing games,
[12] A. Dobrinevski, E. Frey, Extinction in neutrally stable stochastic lotka-volterra models, Phys. Rev. E 85 (2012) 051903.
[13] A. Roman, D. Konrad, M. Pleimling, Cyclic competition of four species: domains and interfaces, Journal of Statistical Mechanics: Theory
[14] P. P. Avelino, D. Bazeia, L. Losano, J. Menezes, von neummann's and related scaling laws in rock-paper-scissors-type games, Phys. Rev. E
[15] P. P. Avelino, D. Bazeia, L. Losano, J. Menezes, B. F. Oliveira, Junctions and spiral patterns in generalized rock-paper-scissors models, Phys.
238105.
Phys. Rev. E 83 (2011) 011917.
and Experiment (2012) P07014.
86 (2012) 031119.
Rev. E 86 (2012) 036112.
Rev. E 87 (2013) 032148.
[16] A. Roman, D. Dasgupta, M. Pleimling, Interplay between partnership formation and competition in generalized may-leonard games, Phys.
[17] J. Vukov, A. Szolnoki, G. Szab´o, Diverging fluctuations in a spatial five-species cyclic dominance game, Phys. Rev. E 88 (2013) 022123.
[18] A. Dobrinevski, M. Alava, T. Reichenbach, E. Frey, Mobility-dependent selection of competing strategy associations, Phys. Rev. E 89 (2014)
012721.
032704.
[19] C. Rulquin, J. J. Arenzon, Globally synchronized oscillations in complex cyclic games, Phys. Rev. E 89 (2014) 032133.
[20] P. P. Avelino, D. Bazeia, L. Losano, J. Menezes, B. F. de Oliveira, Interfaces with internal structures in generalized rock-paper-scissors
models, Phys. Rev. E 89 (2014) 042710.
[21] B. Szczesny, M. Mobilia, A. M. Rucklidge, Characterization of spiraling patterns in spatial rock-paper-scissors games, Phys. Rev. E 90 (2014)
[22] L. Varga, J. Vukov, G. Szab´o, Self-organizing patterns in an evolutionary rock-paper-scissors game for stochastic synchronized strategy
updates, Phys. Rev. E 90 (2014) 042920.
[23] S. Mowlaei, A. Roman, M. Pleimling, Spirals and coarsening patterns in the competition of many species: a complex ginzburglandau
approach, Journal of Physics A: Mathematical and Theoretical 47 (16) (2014) 165001.
[24] C. S. Gokhale, A. Traulsen, Evolutionary multiplayer games, Dynamic Games and Applications 4 (4) (2014) 468–488.
[25] P. P. Avelino, D. Bazeia, J. Menezes, B. de Oliveira, String networks in lotkavolterra competition models, Physics Letters A 378 (4) (2014)
393 – 397.
Physics 17 (11) (2015) 113033.
[26] A. Szolnoki, M. Perc, Vortices determine the dynamics of biodiversity in cyclical interactions with protection spillovers, New Journal of
[27] M. F. Weber, G. Poxleitner, E. Hebisch, E. Frey, M. Opitz, Chemical warfare and survival strategies in bacterial range expansions, Journal of
The Royal Society Interface 11 (96) (2014) 20140172.
[28] D. Groselj, F. Jenko, E. Frey, How turbulence regulates biodiversity in systems with cyclic competition, Phys. Rev. E 91 (2015) 033009.
[29] B. Intoy, M. Pleimling, Synchronization and extinction in cyclic games with mixed strategies, Phys. Rev. E 91 (2015) 052135.
[30] G. Szab´o, K. S. Bod´o, B. Allen, M. A. Nowak, Four classes of interactions for evolutionary games, Phys. Rev. E 92 (2015) 022820.
[31] A. Roman, D. Dasgupta, M. Pleimling, A theoretical approach to understand spatial organization in complex ecologies, Journal of Theoretical
[32] P. P. Avelino, D. Bazeia, R. Menezes, J. C. R. E. Oliveira, Bifurcation and pattern changing with two real scalar fields, Phys. Rev. D 79 (2009)
[33] P. P. Avelino, R. Menezes, J. C. R. E. Oliveira, Unified paradigm for interface dynamics, Phys. Rev. E 83 (2011) 011602. doi:10.1103/
Biology 403 (2016) 10 – 16.
085007.
PhysRevE.83.011602.
[34] E. J. Copeland, R. C. Myers, J. Polchinski, Cosmic F- and D-strings, Journal of High Energy Physics 6 (2004) 013.
[35] M. G. Jackson, N. T. Jones, J. Polchinski, Collisions of cosmic F- and D-strings, Journal of High Energy Physics 10 (2005) 013.
[36] L. Sousa, P. P. Avelino, Probing cosmic superstrings with gravitational waves, Phys. Rev. D 94 (2016) 063529.
10
|
1211.4105 | 1 | 1211 | 2012-11-17T10:01:54 | Simultaneously Propagating Voltage and Pressure Pulses in Lipid Monolayers of pork brain and synthetic lipids | [
"physics.bio-ph"
] | Hydrated interfaces are ubiquitous in biology and appear on all length scales from ions, individual molecules to membranes and cellular networks. In vivo, they comprise a high degree of self-organization and complex entanglement, which limits their experimental accessibility by smearing out the individual phenomenology. The Langmuir technique, however, allows the examination of defined interfaces, whose controllable thermodynamic state enables one to explore the proper state diagrams. Here we demonstrate that voltage and pressure pulses simultaneously propagate along monolayers comprised of either native pork brain or synthetic lipids. The excitation of pulses is conducted by the application of small droplets of acetic acid and monitored subsequently employing timeresolved Wilhelmy plate and Kelvin probe measurements. The isothermal state diagrams of the monolayers for both lateral pressure and surface potential are experimentally recorded, enabling us to predict dynamic voltage pulse amplitudes of 0,1 to 3mV based on the assumption of static mechano-electrical coupling. We show that the underlying physics for such propagating pulses is the same for synthetic (DPPC) and natural extracted (Pork Brain) lipids and that the measured propagation velocities and pulse amplitudes depend on the compressibility of the interface. Given the ubiquitous presence of hydrated interfaces in biology, our experimental findings seem to support a fundamentally new mechanism for the propagation of signals and communication pathways in biology (signaling), which is neither based on protein-protein or receptor-ligand interaction nor on diffusion. | physics.bio-ph | physics | Simultaneously Propagating Voltage and Pressure Pulses in
Lipid Monolayers of pork brain and synthetic lipids
J. Griesbauer1,2, S. Bössinger1,2, A. Wixforth1,3, M.F. Schneider2
1 University of Augsburg, Experimental Physics I, D-86159 Augsburg, Germany
2 Boston University, Dept. of Mechanical Engineering, Boston-Massachusetts, USA
3 Augsburg Center for Innovative Technologies (ACIT) and Center for NanoScience (CeNS)
Abstract
Hydrated interfaces are ubiquitous in biology and appear on all length scales from ions,
individual molecules to membranes and cellular networks. In vivo, they comprise a high
degree of self-organization and complex entanglement, which limits their experimental
accessibility by smearing out the individual phenomenology. The Langmuir technique,
however, allows the examination of defined interfaces, whose controllable thermodynamic
state enables one to explore the proper state diagrams.
Here we demonstrate that voltage and pressure pulses simultaneously propagate along
monolayers comprised of either native pork brain or synthetic lipids. The excitation of pulses
is conducted by the application of small droplets of acetic acid and monitored subsequently
employing time-resolved Wilhelmy plate and Kelvin probe measurements. The isothermal
state diagrams of the monolayers for both lateral pressure and surface potential are
experimentally recorded, enabling us to predict dynamic voltage pulse amplitudes of 0,1 –
3mV based on the assumption of static mechano-electrical coupling.
We show that the underlying physics for such propagating pulses is the same for synthetic
(DPPC) and natural extracted (Pork Brain) lipids and that the measured propagation velocities
and pulse amplitudes depend on the compressibility of the interface. Given the ubiquitous
presence of hydrated interfaces in biology, our experimental findings seem to support a
fundamentally new mechanism for the propagation of signals and communication pathways in
biology (signaling), which is neither based on protein-protein or receptor-ligand interaction
nor on diffusion.
1
Introduction
Lipid bilayers are ubiquitously entangled in biological processes, usually comprising the
impact of changes in external conditions and the exposure to various substances. The static
response of membranes to such conditions was extensively studied from a mechanical (p )
thermal (T) and – although to a lesser extend - electrical (y ) perspective [1–13]. Non-
equilibrium studies on the macroscopic dynamics of lipid membranes, on the other hand, are
only sparsely found, to date [14–19]. The propagation of pulses along interfaces, however,
would add a fundamentally new mechanism to the theory of inter and intra-cellular
communication. Furthermore a proof of the existence of pulses propagating macroscopic
distances would favorably support the idea of nerve pulse propagation in analogy t o sound
[20–24].
We have recently shown that acoustic pulses can indeed propagate over macroscopic distances
in lipid monolayers [25], [26]. Such monolayers can serve as an excellent model system as
they allow direct access to their physical parameters. For instance, a monolayer can be
deposited on a water surface, while controlling the thermodynamic state by compression, pH-
changes or by heating/cooling. At the same time, the response of the monolayer system can be
monitored by measuring, for example, the lateral pressure p with a Wilhelmy plate, giving
A ( ∂ A
∂ π ) , where A is the occupied area of the
access to the lateral compressibility κ = −1
monolayer. Moreover, a Kelvin probe allows simultaneous access to the so called surface
potential y of the lipid monolayer [2].
In this manuscript we study voltage pulses on lipid monolayers, which are accompanied by
acoustic pressure pulses, with respect to the dynamic mechano-electrical coupling in lipid
monolayers. Excited by small acetic acid droplets, the pulses are monitored in both lateral
pressure p and surface potential y and find maximal pressure and potential amplitudes of ~
0.3 mN/m and ~3mV, respectively. We show that static measurements of p and y, on the other
hand, allow the qualitative and quantitative calculation of dynamic pulse shapes and
amplitudes in y using pulse measurements in p. Our experiments demonstrate that both pulse
amplitudes and velocities depend on the compressibility k and thereby on the thermodynamic
state of the monolayer. These results are discussed in the framework of a simple linear
hydrodynamic model, which correctly recovers our results.
2
Materials and Methods
Lipid monolayers of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) or total lipid
extract of pork brain (PBTE) were purchased from Avanti Polar Lipid (USA) and spread from
Chloroform to the air/water-interface of a film balance trough [26], [27] (Fig. 1a). After 10
minutes of evaporation, the lateral pressure and surface potential-area isotherm (p-A and y-A)
are recorded by slow (~ 2,5Å2/(min·molecule)) compression of the film by means of a
moveable barrier. The trough is equipped with two pressure sensors (Wilhelmy plates) and a
Kelvin probe (vibrating capacitor method), which can be read out very rapidly (10000
samples/second, 0.01 mN/m and 0.1 mV resolution). Arriving pressure/potential pulses can
thus be directly monitored by their mechanical and electrical responses, whereas the high
sample rates allow for time resolved measurement and subsequent Fourier transform of the
detected pulse shapes. As only the longitudinal pressure pulses within the monolayers are of
interest, an additional barrier is introduced to exclude spurious effects of unwanted water
waves (Fig. 1a). The actual p- and y-pulse is excited in a separate compartment by sudden
addition of a small amount of acetic acid (~ 3ml) to the monolayer surface. Only pulses able to
travel over macroscopic distances will cause the pressure sensors or the Kelvin probe to
respond. To exclude artifacts by, e.g., unwanted water wave effects, we also performed
reference measurements at which the acetic acid droplets were fused onto a pure water
surface. Within the resolution of our experiments, we were not able to detect any recognizable
pulse response during these reference measurements (see supplementary).
Results
Fig. 1b shows a typical result of a propagating p- and y-pulse. The droplet was deposited
onto a DPPC monolayer at t ~ 1s exciting a pulse. This pulse first arrives at pressure sensor 2,
then at the Kelvin probe and eventually at pressure sensor 1 (Fig. 1a). The time delay between
the two pressure sensor responses and their distance (~ 14,5cm) can be used to directly extract
the propagation velocity of the pulse travelling in the excitation compartmen t along the
separation barrier into the detection compartment [25] . Fig. 1b demonstrates that along with
the measured pressure (p) pulses a nearly identically shaped pulse evolves in the surface
potential y of the monolayer. Indeed, theoretical considerat ions [27], [28] even predict that
all thermodynamic observables change along with the surface pressure (or any other
observable). This behavior is experimentally well reproduced for propagating pulses in p and
3
y suggesting the idea of coupled observables also to be correct for dynamic processes.
To further evaluate the coupling of p and y, their respective isotherms for a DPPC monolayer
are displayed in Fig. 2a, together with the corresponding p-y-curve y(p) (Fig. 2a inset). The
latter was extracted by associating the values of p(A) and y(A) for matching areas A. For a
pulse shape p(t) in the lateral pressure, the function y(p) can be used to predict the
corresponding pulse shape in surface potential by y(t) = y(p(t)). For calculations, the
function y(p) is part-wise approximated linearizations with slopes as they are marked in the
inset of Fig. 2a. Fig. 2b thus presents the predictions for the change in surface potential y(t)
based on the pulse p(t) compared to the actual measurement of y(t). Indeed, both prediction
and measurement agree quantitatively and qualitatively without any fitting adjustment,
rendering the usage of static measurement (inset of Fig. 2a) as valid.
Theoretically a purely adiabatic monolayer pulse, being decoupled from its viscous water sub
phase, may be described by the one dimensional, classical wave equation [26], [29].
2 ∂ 2 v m
∂ 2 v m
∂ x 2 =0 , w h e r e c 0= √1
∂ t 2 −c 0
ρ0 κ S
(1)
Here, vm is the velocity field of the monolayer, r0 its lateral density and kS its adiabatic
compressibility, respectively. The latter will be approximated by the isothermal lateral
compressibility kT , which can be directly extracted from measured p-A-isotherms.
Theoretical evaluations of c0 using kT for a DPPC monolayer yield c0 ~ 50-200 m/s, which is
in contradiction to the pulse propagation velocities of c ~ 1 m/s, experimentally determined
here. One way to account for this discrepancy is to include a coupling of the monolayer lipids
to their water sub phase. In the simplest assumption of direct coupling of the monolayer to the
aqueous sub phase the application of the Stokes-Equations to this system leads to the
following extended wave equation (for a detailed derivation see [25]):
∂ 2 v m
∂ t 2 + 1
ρ0
e i π
4 √η w ρw ω
2 ∂ 2 v m
∂ v m
∂ t −c 0
∂ x 2 =0
(2)
Here, hW and rW represent the water viscosity and density, respectively, while w denotes the
pulse's mean wave frequency. We have chosen this monolayer motivated perspective over a
capillary driven approach, due to the clear dependence of the propagation velocity on the
4
√ ω
ηw ρw
(3)
c= ω
R ( k )
elastic properties of the film. Although our approach matches the data excellently, it should be
noted that in an capillary wave base theory the system can be treated as a free interface with a
adsorbed compressible film, too [14, 15]. We currently discussing such an approach with
collaborators from theory and will report the results properly elsewhere.
However, since in the observed pulse shapes, as shown in Fig. 1b, imply low frequencies of
approximately ~ 1Hz the resulting propagation velocity c in equation (2) can be well
approximated b:
8 )√ 1
= c o s− 1( π
κ S
In Fig. 3a, pulse propagation velocities, which were extracted from the run time of the
pressure signal between the two sensors, are shown as solid lines for a DPPC monolayer at
24°C. Dotted lines represent the calculated results from Equ. (3) using the independently
∂ π )
A ( ∂ A
evaluated isothermal compressibility κ T = −1
T
/ 9 Hz / 18 Hz . Both velocities are plotted as a function of p and reveal a very good
agreement. Considering the general constraint kS < kT
[31], it follows that the directly
measured propagation velocities (from kS) have to be faster than the velocities extracted from
Equ. (3) using kT. Therefore, Fig. 3a implies pulse frequencies of w ~ 1 Hz for p < 10 mN/m
and frequencies of at least w ~ 9 Hz for p > 10 mN/m. Employing a Fourier transform of the
pulse shapes reveals a very similar frequency range.
and three different frequencies w ~ 1 Hz
Identical measurements on cell-extracted lipid (PBTE) monolayers confirm the correlation of
propagation velocity and compressibility as implied by Equ. (3) (Fig. 4a). At the same time it
is remarkable that PBTE monolayers also indicate frequencies of w ~ 1 Hz for p < 10 mN/m
and frequencies of at least w ~ 9 Hz for p > 10 mN/m. This would characterize the pulse
frequencies to be pressure dependent. Apart from this the correlation of surface potential and
lateral pressure isotherms (Fig. 5a) indicate the height of surface potential pulses on PBTE
monolayers. As can be seen at the scale of Fig. 5b the height of PBTE potential pulses is one
order of magnitude smaller than those of DPPC monolayers.
Nevertheless the measurements on PBTE monolayers demonstrate the biological relevance,
showing, that the theory of pulse propagation in synthetic monolayers is also applicable to
biological systems.
For the sake of completeness, the excitability (pulse heights) of DPPC monolayers (Fig. 3b)
5
is shown in Fig. 3b. Similar to the propagation velocities, the peak heights show a dependence
following the compressibility of the monolayer and therefore its thermodynamic state. Indeed,
this behavior even applies to PBTE (see Fig 4b) and therefore utterly underlines the general
coupling of all observables via the thermodynamic state.
Conclusion
Our results demonstrate that voltage-pulses in the low mV range, which are inevitably coupled
to acoustic pressure-pulses, can propagate along lipid monolayers. Thermodynamic couplings
known from static–isothermal experiments are therefore also found under non-equilibrium
conditions. Given the proper detection, we conclude from our results that a corresponding pH
and temperature pulse must exist as well, although we expected the latter to be small. As the
lipid bilayer follows – at least qualitatively - the same physics as the monolayer (under certain
conditions), these results are in support of a thermodynamic foundation of the nervous
impulse [20], [21], [24]. More importantly, they provide a basis to propose a new mechanism
of inter- and intracellular communication in biology (e.g. signaling) in general [25].Wherever
interfaces exist, which are locally in contact to another system (e.g. a bath, a substance, etc.),
a finite probability for (spontaneous or controlled) pulse excitation is expected . A protein, for
instance, embedded in a membrane would “experience” the transient collective changes of the
interface (e.g. compression, electric field, etc.) and react accordingly. Exciting candidates are
enzymes, which are known to exhibit a very strong coupling between activity and interfacial
state [31–33]. Whether the mechanical (p-A) or electrical diagrams of state determine the
coupling between propagating induced state and enzyme activity, however, depends on the
mechanical and electrical properties of the single molecule. In any case, we believe that our
studies predict the propagation of pulses in biological interfaces and suggest a novel
alternative way of communication and signaling between biological entities on scales ranging
from individual enzymes to entire membrane complexes.
. .
6
[1] O. Albrecht, H. Gruler, and E. Sackmann, “Polymorphism of phospholipid monolayers,”
Journal De Physique, vol. 39, no. 3, pp. 301–313, 1978.
[2] G. L. Gaines and others, Insoluble monolayers at liquid-gas interfaces. Interscience
Publishers New York, 1966.
[3] J. Krägel, J. B. Li, R. Miller, M. Bree, G. Kretzschmar, and H. Möhwald, “Surface
viscoelasticity of phospholipid monolayers at the air/water interface,” Colloid &
Polymer Science, vol. 274, no. 12, pp. 1183-1187, Dec. 1996.
[4] H. Beitinger, V. Vogel, D. Möbius, and H. Rahmann, “Surface potentials and electric
dipole moments of ganglioside and phospholipid monolayers: contribution of the polar
headgroup at the water/lipid interface,” Biochimica et Biophysica Acta (BBA) -
Biomembranes, vol. 984, no. 3, pp. 293-300, Sep. 1989.
[5] V. Vogel, “Local surface potentials and electric dipole moments of lipid monolayers:
Contributions of the water/lipid and the lipid/air interfaces,” Journal of Colloid and
Interface Science, vol. 126, no. 2, pp. 408-420, Dec. 1988.
[6] H. Brockman, “Dipole potential of lipid membranes.,” Chemistry and physics of lipids,
vol. 73, no. 1-2, pp. 57-79, Sep. 1994.
[7] X. Wang and Y. Liang, “Formation of Monolayer Lipid Membranes in Water and Ethanol
from Bolaamphiphiles.,” Journal of colloid and interface science, vol. 233, no. 2, pp.
364-366, Jan. 2001.
[8] M. Patra et al., “Under the influence of alcohol: the effect of ethanol and methanol on lipid
bilayers.,” Biophysical journal, vol. 90, no. 4, pp. 1121-35, Feb. 2006.
[9] D. Grigoriev, R. Krustev, R. Miller, and U. Pison, “Effect of Monovalent Ions on the
Monolayers Phase Behavior of the Charged Lipid DPPG,” The Journal of Physical
Chemistry B, vol. 103, no. 6, pp. 1013-1018, Feb. 1999.
[10] A. Lucero, M. R. Rodríguez Niño, A. P. Gunning, V. J. Morris, P. J. Wilde, and J. M.
Rodríguez Patino, “Effect of hydrocarbon chain and pH on structural and topographical
characteristics of phospholipid monolayers.,” The journal of physical chemistry. B, vol.
112, no. 25, pp. 7651-61, Jun. 2008.
[11] K. Nag, J. Perez-Gil, a Cruz, and K. M. Keough, “Fluorescently labeled pulmonary
surfactant protein C in spread phospholipid monolayers.,” Biophysical journal, vol. 71,
no. 1, pp. 246-56, Jul. 1996.
[12] T. Nomura and K. Kurihara, “Effects of changed lipid composition on responses of
liposomes to various odorants: possible mechanism of odor discrimination.,”
Biochemistry, vol. 26, no. 19, pp. 6141-5, Sep. 1987.
[13] N. Koyama and K. Kurihara, “Effect of odorants on lipid monolayers from bovine
olfactory epithelium,” Nature, vol. 236, pp. 402-404, 1972.
[14] J. Lucassen, “Longitudinal waves on visco-elastic surfaces,” Journal of Colloid and
Interface Science, vol. 41, no. 3, pp. 491-498, Dec. 1972.
[15] J. Lucassen, “Longitudinal capillary waves. Part 1. Theory,” Transactions of the Faraday
Society, vol. 64, pp. 2221-2229, 1968.
[16] J. Lucassen, “Longitudinal Capillary Waves. Part 2. Experiments,” Trans. Faraday Soc.,
vol. 64, pp. 2230-2235, 1967.
[17] B. A. Noskov, D. A. Alexandrov, and R. Miller, “Dynamic Surface Elasticity of Micellar
and Nonmicellar Solutions of Dodecyldimethyl Phosphine Oxide. Longitudinal Wave
Study.,” Journal of colloid and interface science, vol. 219, no. 2, pp. 250-259, Nov.
1999.
[18] W. Budach and D. Möbius, “Detection of Longitudinal Waves in Resonance with
7
Capillary Waves at the Air-Water Interface by Energy Transfer,” Thin Solid Films, vol.
178, pp. 61-65, 1989.
[19] M. Suzuki, D. Mobius, and R. Ahuja, “Generation and transmission of a surface pressure
,” ☆ Thin Solid Films, vol. 138, no. 1, pp. 151-156, Apr. 1986.
impulse in monolayers
[20] E. Wilke, “Das Problem der Reizleitung im Nerven vom Standpunkte der Wellenlehre
aus betrachtet,” Pflügers Archiv European Journal of Physiology, vol. 144, no. 1, pp.
35–38, 1912.
[21] T. Heimburg and A. D. Jackson, “On soliton propagation in biomembranes and nerves.,”
Proceedings of the National Academy of Sciences of the United States of America, vol.
102, no. 28, pp. 9790-5, Jul. 2005.
[22] T. Heimburg and A. D. Jackson, “The thermodynamics of general anesthesia.,”
Biophysical journal, vol. 92, no. 9, pp. 3159-65, May 2007.
[23] S. S. L. Andersen, A. D. Jackson, and T. Heimburg, “Towards a thermodynamic theory of
nerve pulse propagation.,” Progress in neurobiology, vol. 88, no. 2, pp. 104-13, Jun.
2009.
[24] K. Kaufmann, Action Potentials. Caruaru Brazil, 1989.
[25] J. Griesbauer, S. Bössinger, A. Wixforth, and M. F. Schneider, “On Propagation of 2D
pressure pulses in lipid monolayers and its possible implications for biology (accepted,
to be published),” Physical Review Letters, 2012.
[26] J. Griesbauer, A. Wixforth, and M. F. Schneider, “Wave propagation in lipid
monolayers.,” Biophysical journal, vol. 97, no. 10, pp. 2710-6, Nov. 2009.
[27] D. Steppich, J. Griesbauer, T. Frommelt, W. Appelt, A. Wixforth, and M. F. Schneider,
“Thermomechanic-electrical coupling in phospholipid monolayers near the critical
point,” Physical Review E, vol. 81, no. 6, pp. 1-5, Jun. 2010.
[28] T. Heimburg, “Mechanical aspects of membrane thermodynamics. Estimation of the
mechanical properties of lipid membranes close to the chain melting transition from
calorimetry.,” Biochimica et biophysica acta, vol. 1415, no. 1, pp. 147-62, Dec. 1998.
[29] L. D. Landau and E. M. Lifshitz, Lehrbuch Der Theoretischen Physik - VII -
Elastizitätstheorie. Verlag Harri Deutsch, 2007.
[30] L. D. Landau and E. M. Lifshitz, Lehrbuch Der Theoretischen Physik - V - Statistische
Physik. Verlag Harri Deutsch, 2008.
[31] H. Sandermann, “Regulation of membrane enzymes by lipids,” Biochim. Biophys. Acta,
vol. 5, pp. 209-237, 1978.
[32] R. Verger and G. H. de Haas, “Interfacial Enzyme Kinetics of Lipolysis,” Annu. Rev.
Biophys. Bioeng., vol. 5, p. 77.117, 1976.
[33] H. M. Seeger, L. Aldrovandi, A. Alessandrini, and P. Facci, “Changes in Single K+
Channel Behaviour Induced by a Lipid Phase Transition,” Biophys. J., vol. 11, pp.
3675-3683, 2010.
8
Figure 1 a) Experimental setup of the film balance used for Monolayer-Pulse-Analysis. The
balance trough is equipped with two Wilhelmy type pressure sensors and a Kelvin probe, such
that both lateral pressure and surface potential can be recorded time-resolved. The additional
barrier separates the excitation from the detection site to ensure the suppression of spurious
water waves. b) Time course of the simultaneous readout of all three sensors after pulse
excitation on a DPPC monolayer (24°C). The signals arrive at the sensors according to Fig. 1
a), such that the pulse travels from pressure sensor 2, to the Kelvin probe, to pressure sensor
1.
Figure 2 a) Isothermal, quasistatic recording of both lateral pressure p and surface potential
y of a DPPC monolayer (24°C). Both isotherms exhibit a flat plateau, indicating the phase
transition from the liquid-expanded to the liquid-condensed phase, whereas the initial
formation (before the liquid-expanded phase) is indicated by the first rises of the surface
potential. In the inset, we correlate surface potential and lateral pressure of the same lipid
areas resulting in a Potential-to-Pressure plot. b) Time course of surface potential (green) and
lateral pressure (red) recorded for a traveling pulse at the detection site. Using the correlation
between pressure and potential as indicated by the inset of Fig. 2 a), the pressure course p(t) is
used to calculate a prediction for the potential y(t) (blue).
Figure 3 a) Propagation velocities of the pulses excited by acetic acid. On the one hand, the
velocities are extracted by the runtime difference between the two pressure sensors (distance
~14,5cm) for different lateral pressures of DPPC monolayers (24°C). On the other hand, the
isothermal compressibility kT, as shown in the inset of Fig 3 b), is used in the model of Equ.
(3) and plotted for three different pulse frequencies of ~1Hz, ~9Hz, ~18Hz. Indeed, the
coincidence of model and measurement indicates frequencies of ~1Hz for low lateral
pressures (< 10 mN/m) and frequencies of ~9Hz for high lateral pressures (> 10 mN/m). b)
Measured pulse amplitudes for different lateral pressures of the DPPC-Monolayer (24°C).
Similar to the propagation velocities the amplitudes follow the phase state of the monolayer
indicated by kT.
Figure 4 a) Propagation velocities of the pulses excited by acetic acid on PBTE-Monolayers
(24°C). Both range of velocities and relation to compressibility demonstrate the same
underlying physics as in figure 3a) showing that this is a general behavior independent of the
respective lipid compositions. b) Measured pulse amplitudes for different lateral pressures of
the PBTE-Monolayer (24°C). As expected, the velocity exhibits neither minima nor maxima,
since the corresponding compressibility kT of PBTE monolayers has no distinct maxima or
minima.
Figure 5 a) Isothermal, quasistatic recording of both lateral pressure p and surface potential
y of a PBTE monolayer (24°C). In contrast to the DPPC isotherms of Figure 2 a) no plateau
region can be observed. In the inset, we again correlate surface potential and lateral pressure
of the same lipid areas resulting in a Potential-to-Pressure plot. b) Time course of lateral
pressure (red) recorded for a traveling pulse at the detection site. Using the correlation
between pressure and potential as indicated by the inset of Fig. 5 a), the pressure course p(t) is
used to calculate a prediction for the potential y(t) (blue).
9
Figure 1
Figure 2
Figure 3
10
Figure 4
Figure 5
11
|
1803.07958 | 2 | 1803 | 2018-07-03T13:35:30 | Sub-picosecond proton tunnelling in deformed DNA hydrogen bonds under an asymmetric double-oscillator model | [
"physics.bio-ph"
] | We present a model of proton tunnelling across DNA hydrogen bonds, compute the characteristic tunnelling time (CTT) from donor to acceptor and discuss its biological implications. The model is a double oscillator characterised by three geometry parameters describing planar deformations of the H bond, and a symmetry parameter representing the energy ratio between ground states in the individual oscillators. If the symmetry parameter takes its maximum value of 1, then we recover a known model which produced CTTs too large to be biologically relevant; but this is reduced by up to 40 orders of magnitude as the symmetry parameter is decreased. We discover that unless the symmetry parameter is close to 1 or 0, the proton's CTT under any planar deformation is guaranteed to be below one picosecond, which is a biologically relevant time-scale. This supports theories of links between proton tunnelling and biological processes such as spontaneous mutation. | physics.bio-ph | physics | Sub-picosecond proton tunnelling in deformed
DNA hydrogen bonds under an asymmetric
double-oscillator model
05/07/2021
J. Luo*
8
1
0
2
l
u
J
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
8
5
9
7
0
.
3
0
8
1
:
v
i
X
r
a
*[email protected]
Department of Mathematical Sciences, Durham University
Durham, DH1 3LE, United Kingdom
Abstract
We present a model of proton tunnelling across DNA hydrogen bonds, compute the charac-
teristic tunnelling time (CTT) from donor to acceptor and discuss its biological implications.
The model is a double oscillator characterised by three geometry parameters describing planar
deformations of the H bond, and a symmetry parameter representing the energy ratio between
ground states in the individual oscillators. We discover that some values of the symmetry
parameter lead to CTTs which are up to 40 orders of magnitude smaller than a previous model
predicted. Indeed, if the symmetry parameter is sufficiently far from its extremal values of 1
or 0, then the proton's CTT under any physically realistic planar deformation is guaranteed
to be below one picosecond, which is a biologically relevant time-scale. This supports theories
of links between proton tunnelling and biological processes such as spontaneous mutation.
1 Introduction
In the DNA double helix, the two strands of nucleobases are held together by hydrogen bonds,
each consisting of a proton being covalently bonded with a donor atom from a donor molecule,
and electrostatically attracted to an acceptor atom from an acceptor molecule [1, 2]. Lowdin
proposed that the proton in an H bond may break away from the donor atom and form a new
covalent bond with the acceptor atom, by the mechanism of quantum tunnelling across the
potential barrier between the donor and acceptor, and that this process may cause spontaneous
mutation [3]. McFadden and Al-Khalili later demonstrated that quantum coherence between
the tunnelling proton and its environment can be maintained for biological time-scales, which
validates modelling the proton's dynamics as being entirely quantum mechanical [4].
In a normal H bond, all atoms in the donor and acceptor molecules are co-planar, and the donor
and acceptor atoms are co-linear with the proton. A planar deformation of the normal H bond
is some combination of translations and rotations in the donor-acceptor molecular plane [5 -- 7].
It has been theorised that planar deformations of the H bond can have significant effects on the
characteristic time-scale of proton tunnelling, and Krasilnikov studied these effects by modelling
the potential in the H bond as a double harmonic oscillator which, when the bond is normal,
is symmetric about the potential barrier [8]. It was found that the characteristic tunnelling
time (CTT) of the proton was extremely sensitive to bond deformation, taking values up to
O(1027)s, which was not a biologically relevant time-scale.
We propose a generalisation to Krasilnikov's model, in which we associate the symmetry of the
double-well potential in a normal H bond with a parameter, γ, whose value equals the energy
ratio between a ground-state proton covalently bonded with the donor and one covalently
bonded with the acceptor. When γ takes its maximum value of 1, we recover Krasilnikov's
model; when 0 < γ < 1, the two local wells in the H bond potential are not equivalent, and
the proton has a preferred equilibrium state near the donor rather than acceptor. We further
encode the planar deformation of the H bond in three other parameters, dx, dy representing
relative shifts between the donor and acceptor, and θ representing the relative rotation, all
of which are defined in detail in Section 2. We then derive an analytical expression for the
proton's CTT. Fixing all other parameters such as proton mass and covalent bond lengths at
values appropriate to DNA H bonds, the CTT is a function of γ, dx, dy and θ. We discover
that moderate values of γ guarantee sub-picosecond proton tunnelling, regardless of bond
deformation. In Section 3, we discuss the biological implications of our results.
2 Model and Results
In this Section, we firstly describe the geometry of an H bond under planar deformation, then
define our double-well potential within this H bond, before solving the Schrodinger equation
under this potential to obtain the proton's wavefunction. From this wavefunction, we derive
the proton's CTT. We make the following assumptions and approximations in our model.
Firstly, we consider only stationary bonds, meaning that the bond is not actively undergoing
deformation whilst proton dynamics is taking place. Secondly, we assume that the lengths and
relative angles of all covalent bonds in the donor and acceptor molecules are unaffected by the
deformation. In other words, we only consider translations and rotations of the donor molecule
as a whole and, independently, of the acceptor molecule as a whole. Finally, even though the
proton's global equilibrium is in a covalent bond with the donor atom, we assume that the
proton can exist with a higher energy in a locally-stable state of being covalently bonded to
1
the acceptor atom. That there are two local potential minima for the proton in the H bond is
the foundation of our double-oscillator model.
D
θ
N2
A
N1
y1
O1
y
O
x1
x
y 2
x 2
2
O
O(cid:48)2
y
d
N(cid:48)2
dx
L
D0
L
D(cid:48)
Figure 1: Geometry of a DNA H bond under planar deformation.
Since the H bond is planar, it suffices to model the potential for the proton as a function of
two spatial dimensions. The geometry of the deformed H bond is shown in Figure 1. Thick
lines marked A and D represent, respectively, the acceptor and donor molecules in a deformed
bond, whilst the dotted line D(cid:48) marks where the donor molecule would be in a normal bond.
(cid:48) marks where the donor
N1 and N2 mark the acceptor and donor atoms, respectively, and N2
atom would be in a normal bond. We set up three Cartesian coordinate systems as follows.
Firstly, centred at O1, where a proton could exist in a covalent bond with N1, we have (x1, y1),
with x1 pointing in the −−−→N1O1 direction. Secondly, centred at O2, where a proton could exist
in a covalent bond with N2, we have (x2, y2), with x2 pointing in the −−−→O2N2 direction. Lastly,
centred at O, the saddle point in the double-well potential of the H bond, whose exact position
along −−−→O1O2 depends upon our potential function, we have (x, y), with x pointing in the −−−→O1O2
(cid:48) marks where O2 would be in a normal bond. The bond geometry is entirely
direction. O2
characterised by 5 parameters, which are marked in Figure 1 as L, D0, dx, dy, θ, and defined as
follows. L is the distance between N1 and O1, which we assume to be the same as the distance
(cid:48), since we have assumed that
between N2 and O2, as well as the distance bweteen N2
(cid:48),
no deformation affects the lengths of covalent bonds. D0 is the distance between O1 and O2
in a normal bond. dx and dy are, respectively, the shifts in the x1 and y1 directions of the
donor molecule from its normal position, so that, for instance, dx < 0 represents a shift of the
donor molecule towards the acceptor molecule. Finally, θ is the anticlockwise angle by which
the donor molecule is rotated from its normal orientation, about the point N2. We emphasise
that the shifts are independent from the rotation, which means that the order in which dx, dy
and θ act on the system does not affect its final configuration.
(cid:48) and O2
By comparing the coordinates of an arbitrary point in the three systems, O1x1y1, O2x2y2 and
Oxy, we write down the following coordinate transformation equations.
x1 = (x + λDθ) cos θ1 − y sin θ1,
y1 = (x + λDθ) sin θ1 + y cos θ1,
x2 =(cid:0)x − (1 − λ)Dθ
y2 =(cid:0)x − (1 − λ)Dθ
(cid:1) cos θ2 − y sin θ2,
(cid:1) sin θ2 + y cos θ2,
2
(1a)
(1b)
(1c)
(1d)
where θ1 is the anticlockwise angle from x1 to x, θ2 is the anticlockwise angle from x2 to x,
Dθ is the distance between O1 and O2 in the deformed bond, and λDθ where 0 < λ < 1 is
the distance between O1 and O in the deformed bond. We express θ1, θ2, Dθ and λ in terms of
L, D0, dx, dy and θ as follows.
θ = 2π + θ1 − θ2,
Dθ cos θ1 = D0 + L + dx − L cos θ,
Dθ sin θ1 = dy − L sin θ,
(cid:113)(cid:2)D0 + dx + L (1 − cos θ)(cid:3)2 +(cid:2)dy − L sin θ(cid:3)2
which imply
Dθ =
Dθ
cos θ1 =
D0 + dx + L (1 − cos θ)
cos θ2 = cos θ1 cos θ + sin θ1 sin θ,
,
sin θ1 =
Dθ
,
dy − L sin θ
sin θ2 = sin θ1 cos θ − cos θ1 sin θ,
(2a)
(2b)
(2c)
(3a)
(3b)
(3c)
and λ is dependent upon the form of the potential function over the (x, y) plane. For our
asymmetric double-oscillator model, we consider a potential function V = V1 + V2, with
(cid:110) U1(x1, y1) := 1
(cid:110) U2(x2, y2) := 1
0
0
V1(x, y) =
V2(x, y) =
2mω2
1
1 + g2y2
1
2mω2
2
2 + g2y2
2
if − ∞ < x < 0, −∞ < y < ∞
otherwise
if 0 ≤ x < ∞, −∞ < y < ∞
otherwise
,
,
(4a)
(4b)
(cid:0)x2
(cid:0)x2
(cid:34)
(cid:1)
(cid:1)
(cid:16)
(cid:35)
(cid:17)
and the proton wavefunction, Ψ(x, y, t), evolves in time according to the Schrodinger equation,
idΨ
dt
=
−
2
2m
∂2
x + ∂2
y
+ V
Ψ.
(5)
In eqs. (4) and (5), m is proton mass, ω1 and ω2 respectively are natural angular frequencies
of the single oscillators U1 and U2, and g > 0 is an isotropy parameter which we assume to be
the same for U1 and U2. We define the symmetry parameter,
γ := ω2/ω1 ≤ 1,
(6)
so that if γ < 1 then there is a lower ground state in U2 than in U1, and this represents the
fact that the proton's preferred equilibrium is in U2. V is a double oscillator which is identical
to U1 to the left of the line x = 0 and identical to U2 to the right of x = 0. Thus, there is a
potential barrier along the line x = 0 where, in general, we have U1 (cid:54)= U2, so that there is a
discontinuity in V .
With the potential function in place, we now calculate λ.
In the O1x1y1 frame, the local
potential well's equipotential curve through the point O is an ellipse, with equation x2
1 =
2U0/(mω2
1), where U0 is the potential energy at O. One could write a similar ellipse equation,
in terms of (x2, y2), for the equipotential curve through O in U2. Instead, using eqs. (1) and (2),
we write both ellipse equations in the Oxy frame, as follows.
1 + g2y2
(cid:2)(x + λDθ) cos θ1 − y sin θ1
(cid:105)2
(cid:1) cos θ2 − y sin θ2
(cid:3)2 + g2(cid:2)(x + λDθ) sin θ1 + y cos θ1
(cid:3)2 =
(cid:105)2
+ g2(cid:104)(cid:0)x − (1 − λ)Dθ
(cid:1) sin θ2 + y cos θ2
=
(cid:104)(cid:0)x − (1 − λ)Dθ
2U0
mω2
1
2U0
mω2
2
,
.
(7a)
(7b)
3
Since the ellipses intersect at O, we set (x, y) = (0, 0) in eqs. (7a) and (7b), to obtain
U0 =
mω2
1
2
λ2D2
θ
from which it follows that
(cid:1) =
(cid:0)cos2 θ1 + g2 sin2 θ1
1 +
(cid:115)
λ =
1
γ
mω2
2
2
(1 − λ)2D2
θ
cos2 θ1 + g2 sin2 θ1
cos2 θ2 + g2 sin2 θ2
(cid:1) ,
(cid:0)cos2 θ2 + g2 sin2 θ2
−1
.
(8)
(9)
We proceed to compute the characteristic time-scale of proton tunnelling from being localised
in U2 to being maximally localised in U1, using the Rayleigh-Ritz ansatz [9], in which the
ground state wavefunction of the proton is approximately
Ψ(x, y, t) = α1(t)φ1(x, y) + α2(t)φ2(x, y),
(10)
where α1,2 are complex coefficients, and φ1,2 are normalised ground state wavefunctions that
the proton would have if it existed in the single-well potential U1 or U2, with their domains
extended to the infinite plane. We note that if a proton were in the single oscillator U1 or U2,
then its ground state energy would be
E1 := ω1(1 + g)/2 for U1
or E2 := ω2(1 + g)/2 for U2,
so that the symmetry parameter, γ, equals the energy ratio E2/E1. Scaling length by
we have φ1 and φ2 in the following dimensionless forms, in terms of coordinates ξ1,2 := x1,2/x0
and η1,2 := y1,2/x0.
(cid:20)
φ1(ξ1, η1) =
φ2(ξ2, η2) =
g1/4
exp
√π
g1/4√γ
√π
−
1
2
(cid:20)
exp
(cid:1)(cid:21)
, −∞ < ξ1, η1 < ∞,
, −∞ < ξ2, η2 < ∞.
Scaling time by ω−1
1 , then Ψ evolves according to the dimensionless Schrodinger equation,
where τ is dimensionless time and, in coordinates (ξ, η) = (x, y)/x0, we have
(cid:16)
(cid:17)
∂2
x + ∂2
y
+ V
= −
1
2
∂2
ξ + ∂2
η
+ v1 + v2,
(15)
(cid:34)
−
1
ω1
2
2m
(cid:16)
(cid:98)H =
(cid:0)ξ2
(cid:110) u1(ξ1, η1) := 1
(cid:0)ξ2
(cid:110) u2(ξ2, η2) := γ2
0
2
2
0
(cid:1)
(cid:1)
1 + g2η2
1
2 + g2η2
2
with
v1(ξ, η) =
v2(ξ, η) =
if − ∞ < ξ < 0, −∞ < η < ∞
otherwise
if 0 ≤ ξ < ∞, −∞ < η < ∞
otherwise
.
,
(16a)
(16b)
4
(cid:114)
x0 :=
,
mω1
(cid:1)(cid:21)
(cid:0)ξ2
γ
2
−
1 + gη2
1
2 + gη2
2
(cid:0)ξ2
= (cid:98)HΨ,
(cid:35)
i
dΨ
dτ
(cid:17)
(11)
(12)
(13a)
(13b)
(14)
Since ∂2
η = ∂2
ξ1,2
+ ∂2
η1,2, we have the following identities.
(cid:18)1
ξ + ∂2
(cid:98)Hφ1 =
(cid:19)
(cid:18) γ
(cid:98)Hφ2 =
(1 + g) − u1 + v1 + v2
φ1,
2
(1 + g) − u2 + v1 + v2
φ2,
(17)
2
(cid:19)
where u1,2 and φ1,2 are expressed in terms of coordinates (ξ, η) as follows. Defining
∆θ := Dθ/x0,
and using the dimensionless version of eq. (1), we obtain, for j = 1, 2,
uj =
1
2
where
(cid:0)ajξ2 + bjη2 + 2cjξη + 2pjξ + 2qjη + rj
(cid:1) ,
a2 = γ2(cid:0)cos2 θ2 + g2 sin2 θ2
(cid:1) ,
b2 = γ2(cid:0)sin2 θ2 + g2 cos2 θ2
(cid:1) ,
c2 = γ2(cid:0)g2 − 1(cid:1) cos θ2 sin θ2,
a1 = cos2 θ1 + g2 sin2 θ1,
b1 = sin2 θ1 + g2 cos2 θ1,
c1 =(cid:0)g2 − 1(cid:1) cos θ1 sin θ1,
We note that λ [cf. eq. (9)] can now be written
p1 = a1λ∆θ,
q1 = c1λ∆θ,
r1 = a1λ2∆2
θ,
p2 = −a2 (1 − λ) ∆θ,
q2 = −c2 (1 − λ) ∆θ,
r2 = a2 (1 − λ)2 ∆2
θ.
(cid:16)
1 +(cid:112)a1/a2
(cid:17)−1
,
λ =
from which it follows that r1 = r2. We therefore define
r0 := r1 = r2 =
(cid:0)√a1 + √a2
a1a2∆2
θ
(cid:1)2 .
For φj with j = 1, 2, we have, for −∞ < ξ, η < ∞,
(cid:20)
(cid:0)Ajξ2 + Bjη2 + 2Cjξη + 2Pjξ + 2Qjη + Rj
(cid:1)(cid:21)
,
φj(ξ, η) =
where
j−1
2 exp
γ
g1/4
√π
1
2
−
A1 = cos2 θ1 + g sin2 θ1, A2 = γ(cid:0)cos2 θ2 + g sin2 θ2
B1 = sin2 θ1 + g cos2 θ1, B2 = γ(cid:0)sin2 θ2 + g cos2 θ2
(cid:1) ,
(cid:1) ,
(18)
(19)
(20a)
(20b)
(20c)
(20d)
(20e)
(20f)
(21)
(22)
(23)
(24a)
(24b)
(24c)
(24d)
C1 = (g − 1) cos θ1 sin θ1, C2 = γ (g − 1) cos θ2 sin θ2,
P1 = A1λ∆θ, P2 = −A2 (1 − λ) ∆θ,
Q1 = C1λ∆θ, Q2 = −C2 (1 − λ) ∆θ,
Defining the inner product (cid:104)fg(cid:105) :=(cid:82) ∞
θ, R2 = A2 (1 − λ)2 ∆2
R1 = A1λ2∆2
θ.
(cid:33)(cid:32)
−∞ dξ(cid:82) ∞
(24f)
−∞ dη f∗g, we take the inner product of eq. (14)
(cid:32)
(cid:33)
(cid:33)(cid:32)
(cid:33)
(cid:32)
(24e)
with (cid:104)φ1 and (cid:104)φ2 respectively to obtain
α1
α2
1 S
S 1
i
,
(25)
H11 H12
H21 H22
α1
α2
=
5
where the overdot denotes differentiation with respect to τ , and
We note that since φ1, φ2 are positve, square normalised functions, and since φ1 (cid:54)≡ φ2, we have
0 < S < 1. Next, using eq. (17), we deduce
(26)
S = (cid:104)φ1φ2(cid:105) , Hjk = (cid:104)φj(cid:98)Hφk(cid:105) .
(cid:32) 1
(cid:33)
γ
(cid:33)
2 (1 + g) + I11
1
2 (1 + g) S + I21
γ
2 (1 + g) S − I12
2 (1 + g) − I22
,
(27)
(cid:82) 0
−∞ dξ(cid:82) ∞
(cid:82) 0
−∞ dξ(cid:82) ∞
−∞ dη (u2 − u1) φ2
−∞ dη (u2 − u1) φ1φ2
1
−∞ dη (u2 − u1) φ1φ2
−∞ dη (u2 − u1) φ2
2
(cid:33)
,
(28)
where(cid:32)
(cid:33)
=
I11
I21
I12
I22
(cid:32)
H11 H12
H21 H22
=
(cid:32) (cid:82) ∞
0 dξ(cid:82) ∞
(cid:82) ∞
0 dξ(cid:82) ∞
(cid:90) ∞
(cid:16)
By invoking the change of variable ξ (cid:55)→ −ξ where necessary, we write, for j = 1, 2 and k = 1, 2,
(cid:90) ∞
(cid:16)
0
(cid:17)
Ajkξ2 + Bjkη2 + 2(−1)k−1Cjkξη + 2(−1)k−1Pjkξ + 2Qjkη + Rjk
aξ2 + bη2 + 2(−1)k−1cξη + 2(−1)k−1pξ + 2qη
−∞
dη
dξ
(cid:17)(cid:21)
√g
2π
j+k
2 −1
γ
(cid:20)
exp
1
2
−
,
(29)
Ijk =
where a = a2 − a1, Ajk = Aj + Ak, and analogous definitions hold for b, Bjk, c, Cjk, p, Pjk, q, Qjk
and Rjk. Each transition integral Ijk can be evaluated exactly, as can the overlap integral, S.
We present closed-form expressions for these integrals in the Appendix.
To solve eq. (25) for αj(τ ), we write(cid:32)
(cid:33)
(cid:32)
(cid:33)(cid:32)
(cid:33)
,
(30)
The solution of eq. (30) subject to the initial condition, (α1, α2) = (0, 1) at τ = 0, is
where
(cid:33)
(cid:32)
J K
M N
= −i
1 − S2
(cid:32) 1
= −i
(cid:16) α1
(cid:17)
α2
ρ± =
J + N ± Ω
2
,
where
with
α1
α2
=
J K
M N
α1
α2
(cid:32)
1 −S
1
−S
(cid:33)(cid:32)
(cid:33)
H11 H12
H21 H22
2 (1 + g) + I11−SI21
1−S2
I21−SI11
1−S2
− I12−SI22
1−S2
γ
2 (1 + g) − I22−SI12
1−S2
(cid:33)
.
(β+r+eτ ρ+ + β−r−eτ ρ−) ,
=
1
Nτ
r± =
(cid:18)
1, −J + N ± Ω
(cid:19)T
Ω =(cid:112)(J − N )2 + 4KM ,
2K
6
,
β± = ±K/Ω,
(31)
(32)
(33)
(34)
and we determine the real function Nτ as follows. From eq. (32), we have
(cid:18) J + N
(cid:18) J + N
2
(cid:19)
(cid:19)(cid:20)
τ
sinh
,
Ωτ
2
Ωτ
2 −
τ
cosh
2
α1 =
α2 =
exp
2K
Nτ Ω
1
Nτ
exp
(cid:21)
(35a)
(35b)
.
(J − N )
Ω
sinh
Ωτ
2
Assume for now that Ω2 < 0, which we later verify numerically, so that Ω = iΩ, then we have
(36)
cosh
sinh
,
.
Ωτ
2
= i sin Ωτ
2
Ωτ
2
= cos Ωτ
2
Since e(J+N )τ /2 = 1, it follows from the normalisation condition, (cid:104)ΨΨ(cid:105) = α12 + α22 +
(α∗
2α1)S = 1, that
1α2 + α∗
(cid:114)
Nτ =
cos2 Ωτ
2
+ σ sin2 Ωτ
2
,
(37)
where σ = (4K2 + J − N2 + 4K(J − N )S)/Ω2, which is real because K(J − N ) is real.
Since S < 1, we have σ > (4K2 + J − N2 − 4K(J − N ))/Ω2 = (2K − J − N)2/Ω2,
therefore σ > 0. In the proton wavefunction Ψ = α1φ1 + α2φ2, α2 is initially unity and α1 is
initially zero, so we say that the proton's CTT, the time it takes for Ψ to evolve from being
localised as φ2 to being maximally localised in the potential well u1, is the time at which
(cid:12)(cid:12)(cid:12)(cid:12)sin Ωτ
2
(cid:12)(cid:12)(cid:12)(cid:12)
α1 =
2K
NτΩ
(cid:18)
(cid:19)
(38)
sin Ωτ
2
(39)
(40)
first reaches its maximum. This happens at the smallest τ for which the following holds.
0 =
d
dτ
Ωτ
2
sin
Nτ
= Ω
2Nτ
= Ω
2N 3
τ
cos Ωτ
2
cos Ωτ
2
+ Ω
2N 3
τ
.
cos Ωτ
2
sin Ωτ
2 − σ sin Ωτ
2
cos Ωτ
2
Therefore, the CTT of the proton is τp = π/Ω, or, in physical units,
tp =
π
ω1Ω
,
(cid:115)
(cid:20)1
2
Ω =
−
where, due to eqs. (31) and (34), we have
(1 + g) (1 − γ) +
I11 + I22 − S (I12 + I21)
1 − S2
(cid:21)2
+
4 (I12 − SI22) (I21 − SI11)
(1 − S2)2
.
(41)
We have the following values for the parameters D0, L and g which are appropriate for H
bonds across the DNA double helix [10 -- 13]. 4.5 × 1014s−1 ≤ ω1 ≤ 6.4 × 1014s−1, 0.61A ≤ D0 ≤
0.81A, 1.03A ≤ L ≤ 1.07A, g ≈ 0.5. We fix ω1 = 5.45 × 1014s−1, D0 = 0.71A, L = 1.05A, g =
0.5, and compute tp as functions of the parameters γ, dx, dy and θ. For all parameter values
which we have studied, we find Ω2 < 0, which ensures that eq. (36) holds. We note also that
when γ = 1, we recover results of [8] relating to deformations of a symmetric double oscillator.
In order for our model to represent tunnelling, rather than scattering, we must have the height
U0 [cf. eq. (8)] of the saddle point in the double-well potential surface being greater than the
ground-state energy of φ2 [cf. eq. (11)]; that is, we must have
u0 :=
U0
E2
=
r0
γ(1 + g)
> 1.
7
(42)
Figure 2: dcrit
x
as a function of γ.
Figure 3: Min, max, average tp as functions of γ.
Moreover, the expected value of proton energy must be conserved by the tunnelling process;
that is, we must have d(cid:104)Ψ(cid:98)HΨ(cid:105) /dτ = 0. Since (cid:98)H is time-independent, we do indeed have
d(cid:104)Ψ(cid:98)HΨ(cid:105) /dτ = (cid:104) Ψ(cid:98)HΨ(cid:105) + (cid:104)Ψ(cid:98)H Ψ(cid:105) = i(cid:104)Ψ(cid:98)H(cid:98)HΨ(cid:105) − i(cid:104)Ψ(cid:98)H(cid:98)HΨ(cid:105) = 0, where we have made
use of the Schrodinger equation and its dual, −i(cid:104) Ψ = (cid:104)Ψ (cid:98)H. We note that the proton
wavefunction for τ > 0 is always a superposition of φ1 and φ2 with a non-zero coefficient for
φ2 [cf. eq. (35)], for if that coefficient were to vanish at any time then the proton energy at
that time would equal E1 > E2, violating the energy conservation requirement.
x
x
then eq. (42) is satisfied given any combination of (dy, θ), whereas if dx < dcrit
The deformation parameters dx, dy and θ are encoded in r0, as per the definition of eq. (22).
Our results show that, for each value of γ, there exists some critical value dcrit
such that, if
dx ≥ dcrit
then
there are some combinations of (dy, θ) under which eq. (42) fails to hold. Figure 2 shows dcrit
as a function of γ. As γ decreases towards 0, greater values of dx would be needed in order to
guarantee that every combination of (dy, θ) produces a valid tunnelling model. This is because
γ is positively correlated with the steepness of the local potential well U2(x2, y2). The smaller
γ is, the further away from (x2, y2) = (0, 0) one needs to go before U2 reaches the required
height, namely the ground-state energy of φ2; thus, in order to ensure that the saddle point
between U1 and U2 is sufficiently high, U1 and U2 must be far enough apart, hence the large
dcrit
x . Meanwhile, as γ → 1, we observe that dcrit
For 0.01 ≤ γ ≤ 1, we vary dx, dy, θ as follows. −0.45A ≤ dx, dy ≤ 0.45A,−90◦
we only consider combinations of (γ, dx, dy, θ) such that eq. (42) holds. We find that for each
γ, tp falls in a range between some tmin
(γ), and in Figure 3 we present these
extremal values as functions of γ. Crucially, our results show that for 0.01 ≤ γ ≤ 0.99, we
always have 8.5fs ≤ tp(γ, dx, dy, θ) ≤ 770fs. We also observe that tmax
(γ) increases steeply both
as γ → 0 and as γ → 1. Indeed, when γ = 1, tmax
(γ) becomes ∼ O(1027)s; and even though
(γ) is still ∼ O(10−14)s, tp increases rapidly as (dx, dy, θ) moves away from the combination
tmin
p
which minimises tp. Moreover, tmin
(γ) is slowly varying with γ, and there is a range of values
of γ, namely 0.2 (cid:47) γ (cid:47) 0.4, for which tmin
(γ). In this case, varying
(dx, dy, θ) has little effect on tp, which contrasts strongly with the large-γ and small-γ cases
where tp is very sensitive to (dx, dy, θ). We have defined tave
p (γ) as the mean tp, given a fixed
γ, over all combinations of (dx, dy, θ) which satisfy eq. (42), and we have presented tave
p (γ) for
0.1 ≤ γ ≤ 0.6 in the small box in Figure 3. As γ → 1, we have tave
(γ), and for
intermediate values of γ, namely γ ≈ 0.3, we have tave
p (γ) is
asymptotic to neither tmin
(γ), but as γ → 0, tave
≤ θ ≤ 90◦, and
x → −0.44A.
(γ) becomes close to tmax
p (γ) ∼ tmax
p
p (γ) ∼ tmin
p
(γ) nor tmax
(γ).
p
p
(γ) and some tmax
p
p
p
p
p
p
x
x
p
8
0.20.40.60.81−0.6−0.300.30.6γdcritx(A)0.20.40.60.8104080120160γtp(fs)0.20.30.40.58121620 tminp(γ)tmaxp(γ)tavep(γ)(a) γ = 0.55.
(b) γ = 0.85.
Figure 4: tp as functions of θ, given various combinations of (γ, dx, dy).
Furthermore, our results show that for every (γ, dx), we have
tp(γ, dx, dy, θ) = tp(γ, dx,−dy,−θ).
(43)
This is because a deformation consisting of a shift of dy and rotation of θ is intrinsically
identical to one consisting of a shift and rotation of the same magnitudes but both in the
opposite direction. Figure 4 shows variations in tp as θ varies between −90◦ and 90◦, whilst
(γ, dx, dy) are fixed at certain values. For every combination of (γ, dx), we have presented only
results relating to dy ≥ 0, since one can simply reflect these curves about θ = 0 to obtain results
for dy < 0. For fixed (γ, dx) with dy = 0, the graph of tp(θ) is symmetric about θ = 0, where
the graph has a local mimimum under some (γ, dx) and a local maximum under others; we find
from our results that for every γ there is one value of dx at which the graph transitions from
having a local minimum to having a local maximum at θ = 0, and that this value of dx increases
with γ. For fixed (γ, dx) with dy (cid:54)= 0, the symmetry of tp(θ) about θ = 0 is broken, and as dy
increases, the local extremum which was at θ = 0 when dy = 0 moves towards larger θ. There
are cases where this local extremum ceases to exist when dy becomes large, for instance the
9
−60−300306017.0317.41 dx=−0.3A,dy=0Adx=−0.3A,dy=0.2Adx=−0.3A,dy=0.4A−60−300306017.06217.098tp(fs) dx=0A,dy=0Adx=0A,dy=0.2Adx=0A,dy=0.4A−60−300306017.0793417.07973θ(degrees) dx=0.3A,dy=0Adx=0.3A,dy=0.2Adx=0.3A,dy=0.4A−60−30030604452 dx=−0.3A,dy=0Adx=−0.3A,dy=0.2Adx=−0.3A,dy=0.4A−60−300306051.23251.242tp(fs) dx=0A,dy=0Adx=0A,dy=0.2Adx=0A,dy=0.4A−60−300306051.239021751.2390242θ(degrees) dx=0.3A,dy=0Adx=0.3A,dy=0.2Adx=0.3A,dy=0.4Acase of (γ, dx) = (0.55,−0.3A), as we can see in Figure 4a: there is a local minimum at θ = 0 if
dy = 0 and at θ = 5◦ if dy = 0.2A, but if dy = 0.4A then this local mimimum disappears. For
any fixed (γ, dx, dy), we always have tp tending to some value as θ tends to ±90◦, typically with
several local extrema between θ = 0 and θ = ±90◦; the value of this limit at ±90◦ is dependent
only on γ. Calling this limit t90
p (0.85) = 51.2fs. As γ → 1
and as γ → 0, we have t90
(γ), and for 0.02 (cid:47) γ (cid:47) 0.4, we have t90
p (0.55) = 17.1fs, and t90
p (γ), we have t90
p (γ) ∼ tmax
p
p (γ) ∼ tmin
p
(γ).
(a) γ = 0.25, θ = 0.
(b) γ = 0.55, θ = 0.
(c) γ = 0.85, θ = 0.
(d) γ = 0.25, θ = 40◦.
(e) γ = 0.55, θ = 40◦.
(f) γ = 0.85, θ = 40◦.
Figure 5: tp as surfaces over the parameter subspace (dx, dy), given various
combinations of (γ, θ). In each case, the range of dx is dcrit
x ≤ dx ≤ 0.45A.
p
(γ) and tmax
We further observe by comparing Figures 4a and 4b that, when γ = 0.85, there is a larger
overall variation in tp as a result of varying (dx, dy, θ), compared to when γ = 0.55. This agrees
with our observation about Figure 3 that the gap between tmin
(γ) increases as
γ → 1. Indeed, this gap also increases as γ → 0. Moreover, for fixed γ, the larger dx is, the
less tp varies with θ or with dy. As we see in Figures 5b, 5c, 5e and 5f, if γ is far from 0, then
for fixed (γ, θ), tp as a surface over (dx, dy) is almost constant given sufficiently large dx. As
dx → ∞, tp always tends to some limit, whose value is independent of dy. Meanwhile, we see
in Figures 5a to 5c that if θ = 0, then for fixed (γ, θ), tp as a surface over (dx, dy) is symmetric
about the line dy = 0. This is due to eq. (43). If θ = 0 and γ is moderate, such as 0.55, then
for each dy sufficiently to 0 we have some small value of dx which maximises tp, as we can see
in Figure 5b. This shows that increasing dx, which represents moving the donor away from
the acceptor in the H bond, does not necessarily prolong the proton tunnelling. If θ (cid:54)= 0, then
the symmetry about dy = 0 is broken, and reflecting a surface for θ > 0 about the line dy = 0
produces corresponding results for θ < 0.
p
10
−0.50−0.2500.250.50−0.50−0.2500.250.501010.51111.51212.5dx(A)dy(A)tp(fs)−0.50−0.2500.250.50−0.50−0.2500.250.501415161718dx(A)dy(A)tp(fs)−0.50−0.2500.250.50−0.50−0.2500.250.50102030405060dx(A)dy(A)tp(fs)−0.50−0.2500.250.50−0.50−0.2500.250.501010.51111.5dx(A)dy(A)tp(fs)−0.50−0.2500.250.50−0.50−0.2500.250.501717.117.217.317.4dx(A)dy(A)tp(fs)−0.50−0.2500.250.50−0.50−0.2500.250.504949.55050.55151.5dx(A)dy(A)tp(fs)3 Discussions and Conclusions
We have studied the quantum mechanical tunnelling of a proton across the potential barrier
between the donor and acceptor of a planar hydrogen bond in DNA, and computed an ana-
lytical expression for the proton's characteristic tunnelling time (CTT) as a function of four
parameters describing the geometry of the bond. Three of these parameters, dx, dy and θ,
represent the deformation of the H bond from its normal alignment, under the assumption
that any deformation consists of planar translations and rotations of the donor and acceptor
molecules as independent units. With the acceptor molecule treated without loss of generality
as fixed, dx and dy respectively represent the longitudinal and lateral displacements of the
donor molecule from its normal position, while θ represents the rotation of the donor molecule
about the donor atom from its normal orientation. The fourth parameter, γ, taking values
0 < γ ≤ 1, represents the intrinsic symmetry that the potential in the H bond possesses when
the bond is in its normal alignment. When γ = 1, we recover a model previously studied in [8],
whose potential function in the normal H bond was symmetric about the potential barrier, so
that the local potential wells near the donor and acceptor are equivalent to each other. This
symmetry is broken only if some of (dx, dy, θ) is non-zero. For 0 < γ < 1, the symmetry is bro-
ken even if dx = dy = θ = 0, in the sense that the local potential well near the donor has a less
energetic ground state than the one near the acceptor, and this gives a better representation
of the physical property of the H bond than γ = 1. In addition, setting any of dx, dy and θ to
non-zero values further distorts the symmetry between the two local potential wells.
p
(γ) or tmax
p
p
p
p
p (γ) ∼ tmax
p
We have discovered that some combinations of (γ, dx, dy, θ) provide potential functions which
cannot model a tunnelling process, because the potential barrier is not higher than the ground
state energy of a proton in equilibrium near the donor. The smaller γ is, the more (dx, dy, θ)
combinations provide invalid models, meaning that the region of validity in our parameter space
shrinks as γ decreases. For 0.01 ≤ γ ≤ 0.99,−0.45A ≤ dx, dy ≤ 0.45A,−90◦
≤ θ ≤ 90◦, and ex-
cluding all invalid parameter combinations, we have found that 8.5fs ≤ tp(γ, dx, dy, θ) ≤ 770fs,
where tp stands for the proton's CTT. For each γ, certain (dx, dy, θ) combinations minimise or
maximise tp, and we have found that tmin
(γ) is a slowly-varying function taking values around
(γ) diverges as γ → 0 and grows rapidly towards O(1027)s as γ → 1. Taking
10fs, whilst tmax
the mean tp over all (dx, dy, θ) for every fixed γ, we have found that tave
(γ) as γ → 1.
This means that in an H bond selected at random from a statistical ensemble, the proton's
CTT is likely to be as large as it can be if the potential in the bond has a high γ-symmetry.
On the other hand, we have also observed that if γ takes moderate values such as γ ≈ 0.3,
then tave
(γ), meaning that the proton's CTT is likely to be as small as it can be in
this case. As γ → 0, tave
(γ)
diverges towards infinity in this case, we deduce that parameter combinations resulting in large
tp are rare when γ is small. We have investigated how tp varies with θ given fixed (γ, dx, dy),
and found that as θ → ±90◦, tp always converges to some t90
p (γ) which depends on γ in the
following manner. In extreme cases of γ → 1 and γ → 0, we have t90
(γ), and for
(γ). For −90◦ < θ < 90◦, we have observed that tp has
moderate γ values, we have t90
various local maxima and local minima but the variation in tp is small unless either γ is close
to extremal values, or dx is negative with large magnitudes. For example, if 0.3 ≤ γ ≤ 0.99
and dx ≥ 0, then regardless of dy, we have the result that as θ varies, tp never deviates by more
than 1% from some average value. We have also investigated how tp varies with (dx, dy), given
fixed (γ, θ), and found that if dx is sufficiently large, then tp is an almost-constant surface over
(dx, dy), and that tp tends to some dy-independent limit as dx → ∞. Since large dx corresponds
to large donor-acceptor separation, one might expect tp to be maximised in the limit dx → ∞,
but our results show that this is not always the case.
p (γ) is not asymptotic to tmin
(γ); given the fact that tmax
p (γ) ∼ tmax
p
p (γ) ∼ tmin
p
p (γ) ∼ tmin
p
11
The most important difference that generalising from γ = 1 to 0 < γ ≤ 1 has made is that,
for most γ values in 0 < γ < 1, the proton CTT is sub-picosecond regardless of (dx, dy, θ).
Compared to the γ = 1 case in which some (dx, dy, θ) give CTTs of O(1027)s, the sub-picosecond
time-scale is much more biologically relevant. Moreover, if γ is such that the CTT is guaranteed
to be sub-picosecond, then it varies by no more than 2 orders of magnitude as the H bond
deforms. This means that the tunnelling process is much more stable with respect to bond
deformation compared to the γ = 1 case, under which the CTT varies by over 30 orders of
magnitude as the H bond deforms. Overall, our model under moderate γ-values produces
CTTs on a biological time-scale with strong stability against bond deformation, and therefore
it supports the theory that proton tunnelling across DNA hydrogen bonds may be a mechanism
responsible for biological processes such as spontaneous mutation.
The author is grateful to Dr. Emma Coutts and Dr. Bernard Piette for their kind support.
Appendix
In Section 2 we presented the overlap integral S and transition integrals Ijk, for j, k = 1, 2 [cf.
eqs. (26) and (29)]. We have computed closed-form expressions for these integrals, as follows.
S = 2
√gγ
K0,12
exp
Ijk = √gγ
j+k
2 −1
(cid:35)
(cid:33)
,
K 2
2,12
2B12K 2
0,12
b − 2qQjk
Bjk
(cid:32)
K1,12 +
(cid:34)
(cid:32)
+ 2(−1)k−1
(cid:32)
bQ2
jk
B2
jk
+
+
bC 2
jk
jk −
B2
(cid:33)
where
1
K 2
exp
2K0,jk
J1,jk =
J0,jk =
K1,jk +
2,jk
2BjkK 2
(cid:112)Bjk
(cid:32)
exp(cid:0)K1,jk
(cid:1)
(cid:32)
√2πK 2
+ (−1)k K2,jk
(cid:112)BjkK2,jk
exp(cid:0)K1,jk
(cid:1)
(cid:16)
(cid:17)
(cid:32)
J2,jk = (−1)k
K 2
√2πK 4
2,jk + BjkK 2
K1,jk +
2K 3
exp
0,jk
0,jk
0,jk
0,jk
(44a)
(44b)
(45a)
(45b)
(cid:33)
,
(45c)
(cid:1)
(cid:33)
+ p
J1,jk
(cid:33)
,
J0,jk
(cid:0)cQjk + qCjk
,
Bjk
bCjkQjk
(cid:33)
jk −
B2
2cCjk
Bjk
+ a
J2,jk
(cid:32)
K2,jk
(−1)k−1(cid:112)2BjkK0,jk
(cid:32)
(cid:33)
(cid:33)
,
K2,jk
(−1)k−1(cid:112)2BjkK0,jk
(cid:32)
(cid:33)
(−1)k−1(cid:112)2BjkK0,jk
K2,jk
erfc
0,jk
K 2
2,jk
2BjkK 2
0,jk
erfc
K 2
2,jk
2BjkK 2
0,jk
12
+
2K 5
0,jk
exp
K1,jk +
erfc
with
K0,jk =
(cid:113)
AjkBjk − C 2
jk, K1,jk =
, K2,jk = BjkPjk − CjkQjk,
(46)
Q2
jk
2Bjk −
Rjk
2
(cid:90) ∞
and erfc being the cumulative error function, defined for all real X by
erfc(X) = (2/√π)
e−z2dz.
(47)
The parameters a, b, c, p, q, Ajk, Bjk, Cjk, Pjk, Qjk, Rjk were defined in the main text.
X
References
[1] L. Pauling. The Nature of the Chemical Bond. Cornell University Press, 3rd edition, 1960.
[2] E. Arunan, G. R. Desiraju, R. A. Klein, J. Sadlej, S. Scheiner, I. Alkorta, D. C. Clary,
R. H. Crabtree, J. J. Dannenberg, P. Hobza, H. G. Kjaergaard, A. C. Legon, B. Mennucci,
and D. J. Nesbitt. Pure Appl. Chem., 83:1637, 2011.
[3] P.-O. Lowdin. Rev. Mod. Phys., 35:724, 1963.
[4] J. McFadden and J. Al-Khalili. BioSystems, 50:203, 1999.
[5] R. E. Dickerson. Nucleic Acids Research, 17:1797, 1989.
[6] X.-J. Lu and Wilma. K. Olson. J Mol. Biol., 285:1563, 1999.
[7] W. K. Olson, M. Bansal, S. K. Burley, R. E. Dickerson, M. Gerstein, S. C. Harvey,
U. Heinemann, X.-J. Lu, S. Neidle, Z. Shakked, H. Sklenar, M. Suzuki, C.-S. Tung,
E. Westhof, C. Wolberger, and H. M. Berman. J. Mol. Biol., 313:229, 2001.
[8] P. M. Krasilnikov. Biophysics, 59:189, 2014.
[9] E. Merzbacher. Quantum Mechanics. Wiley, 3rd edition, 1998.
[10] S. Ia. Ishenko, M. V. Vener, and V. M. Mamaev. Theor. Chim. Acta, 68:351, 1985.
[11] R. Santamaria, E. Charro, A. Zacar´ıas, and M. Castro. J. Comput. Chem., 20:511, 1999.
[12] C. Fonseca Guerra, F. M. Bickelhaupt, J. G. Snijders, and E. J. Baerends. J. Am. Chem.
Soc., 122:4117, 2000.
[13] T. Steiner. Angew. Chem. Int. Ed., 41:48, 2002.
13
|
1707.02425 | 1 | 1707 | 2017-07-08T10:53:25 | Nonlinear dynamics of damped DNA systems with long-range interactions | [
"physics.bio-ph"
] | We investigate the nonlinear dynamics of a damped Peyrard-Bishop DNA model taking into account long-range interactions with distance dependence |l|^-s on the elastic coupling constant between different DNA base pairs. Considering both Stokes and long-range hydrodynamical damping forces, we use the discrete difference operator technique and show in the short wavelength modes that the lattice equation can be governed by the complex Ginzburg-Landau equation. We found analytically that the technique leads to the correct expression for the breather soliton parameters. We found that the viscosity makes the amplitude of the breather to damp out. We compare the approximate analytic results with numerical simulations for the value s = 3 (dipole-dipole interactions). | physics.bio-ph | physics | Nonlinear dynamics of damped DNA systems with long-range interactions
J. Brizar Okalya,c,∗, Alain Mvogoa,c, R. Laure Woulach´eb,c, T. Cr´epin Kofan´eb,c
aLaboratory of Biophysics, Department of Physics, Faculty of Science, University of Yaounde I, P.O. Box 812, Yaounde, Cameroon
bLaboratory of Mechanics, Department of Physics, Faculty of Science, University of Yaounde I, P.O. Box 812, Yaounde, Cameroon
cAfrican Center of Excellence in Information and Communication Technologies, University of Yaounde I, P.O. Box 812, Yaounde,
Cameroon.
7
1
0
2
l
u
J
8
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
5
2
4
2
0
.
7
0
7
1
:
v
i
X
r
a
Abstract
We investigate the nonlinear dynamics of a damped Peyrard-Bishop DNA model taking into account long-range interactions
with distance dependence l−s on the elastic coupling constant between different DNA base pairs. Considering both Stokes
and long-range hydrodynamical damping forces, we use the discrete difference operator technique and show in the short
wavelength modes that the lattice equation can be governed by the complex Ginzburg-Landau equation. We found
analytically that the technique leads to the correct expression for the breather soliton parameters. We found that the
viscosity makes the amplitude of the breather to damp out. We compare the approximate analytic results with numerical
simulations for the value s = 3 (dipole-dipole interactions).
Keywords: DNA, long-range interactions, damping forces, breather soliton.
1.
Introduction
The DNA molecule is known to be very important and essential in the protection, transport and transmission of the
genetic code. Many theoretical models have been proposed to describe the nonlinear dynamics of DNA. The first nonlinear
DNA model was suggested by Englander et al. [1]. Thereafter, simplified models were proposed to describe the angular
distortion of DNA [2, 3, 4] and the micromanipulation experiments were investigated, showing the great importance of
radial displacements of bases during the processes of replication and transcription [5, 6, 7]. The Peyrard-Bishop (PB) DNA
model [5], which has been successfully used to analyze experiments on short DNA sequences [8] has gained a popularity
in that direction.
DNA double helix spontaneously denatures locally and the breathing mode occurs when it is locally excited with large
amplitude [9]. The amplitude of this excitation depends on the surrounding environment-DNA interactions and the inner
mechanism of the DNA molecule depending on the relative motion of particles. It is therefore of physical importance to
take into account these two effects in the nonlinear dynamics of DNA molecule. Some authors emphasized the influence
of the viscosity on the dynamical properties of the DNA molecule [10, 11]. In particular, it is shown that the viscosity
of the medium damps out the amplitude of the nonlinear wave propagating through the molecule [10, 11]. In the above
studies, the analysis deal with the PB DNA model with nearest neighbor interactions between base pairs. The long-range
interactions (LRI) play a crucial role in molecular systems [12, 13, 14]. The importance of LRI in DNA molecule is due
to the presence of phosphate groups along the strands [12]. The LRI therefore allow to take into account the screening
∗Corresponding author
Email addresses: [email protected] (J. Brizar Okaly), [email protected] (Alain Mvogo), [email protected] (R. Laure
Woulach´e), [email protected] (T. Cr´epin Kofan´e)
Preprint submitted to :
September 11, 2018
of the interactions or an indirect coupling between base pairs (e.g. via water filaments) [13]. Rau and Parsegian, in their
experimental studies in the direct measurements of the intermolecular forces between counterion-condensed DNA double
helices, have shown the importance of long-range attractive hydrogen forces and emphasize that many forces could be
responsible for LRI in the DNA molecule [14]. In fact, the charged groups in the molecular chains and DNA molecules
interact through long-range dipole-dipole interactions. Along the same line, it has been shown that the study of the
power-law LRI in nonlinear lattice models is very relevant mainly in molecular chains and DNA molecules where Coulomb
and dipole-dipole interactions are of a great physical importance [15, 16, 17]. Recent works by Mvogo et al. [18, 19] also
indicated qualitative effects of the power-law LRI in molecular systems.
To the best of our knowledge, no work has been reported on the study of DNA dynamics taking into account both
damping and LRI effects. Our aim in this paper is to study the DNA dynamics taking into account LRI between different
DNA base pairs and both Stokes and long-range hydrodynamical damping effects. To address this issue, our analytical
study has been inspired by the one recently developed by Miloshevich et al.
[20] to investigate traveling solitons in
long-range oscillator chains. Due to the non analytical properties of the dispersion relation, Miloshevich et al. [20] have
shown that the discrete difference operator (DDO) technique is more appropriate to study physical systems with LRI. In
this paper, we use the DDO and show that the DNA models with LRI can be reduced to a specific form of the complex
Ginzburg-Landau (CGL) equation, where the dispersion coefficient is complex and the nonlinearity coefficient is real. A
similar equation has been obtained by Zdravkov´ıc et al. [10, 11] while studying the effect of viscosity on the dynamics of the
Peyrard-Bishop-Dauxois (PBD) DNA model in the absence of hydrodynamical damping and LRI forces. The investigators
state that the CGL equation cannot be solved analytically like a nonlinear Schrodinger (NLS) equation [10, 11]. In this
paper, following the work by Pereira and Stenflo [21], we analytically solve the CGL equation.
The paper is organized as follows. In Section 2, we propose the Hamiltonian model and derive the discrete equations
of motion for the in-of-phase and out-of-phase motions. In Section 3, we use the DDO and show that the out-of-phase
dynamical equation can be reduced to the CGL equation. The envelope soliton solution of this equation is reported and
the breather solution of the discrete equation of motion for the out-of-phase motion is derived. In Section 4, we perform
the numerical simulations with emphasis on the effects of the LRI and damping forces. Section 5 concludes the work.
2. Model and equations of motion
We consider the PB model [5] for DNA denaturation where the degrees of freedom xn and yn associated to each base
pair correspond to the displacements of the bases from their equilibrium positions along the direction of the hydrogen
bonds that connect the two bases in a pair. A LRI coupling between the base pairs due to the presence of phosphate
group along the DNA strands is assumed so that the Hamiltonian for the model is given by
H =
N
Xn n 1
2
m( x2
n + y2
n) +
1
2Xl=1
Jl[(xn − xn−l)2 + (yn − yn−l)2] + V (xn, yn)o,
(1)
where m is the average mass of the nucleotides and N represents the number of the base pairs of the DNA molecule. The
interactions between hydrogen bonds in a pair is modeled by the Morse potential V (xn, yn) given by
V (xn, yn) = Dhe−a(xn−yn) − 1i2
,
2
(2)
where D is the depth of the Morse potential well, which may depend on the type of base pair and a is the width of the
well. The quantity
Jl = Jl−s,
(3)
is the power-law dependence of the elastic coupling constant, where s and l are the LRI parameter and the normalized
distance between base pairs, respectively. In practice, to keep the spatial homogeneity in a finite DNA system with periodic
boundary conditions, usually the LRI is limited in each direction to 1
2 (N − 2), if N is even,
and 1 ≤ l ≤ 1
2 (N − 1) [22]. The parameter s can be used to model Coulomb interactions between charged particles of a
chain (s = 1), dipole-dipole interactions (s = 3). Below s = 1 the energy diverges and above s = 3 the system becomes
2 (N − 1), if N is odd, or 1
short-range. In this paper, we use the case s = 3, where multiple solutions exist [23].
The values of parameters used to perform our analysis are those from the dynamical and denaturation properties of
DNA. They are [24]: m = 300 amu, J = 0.06 eV/A2, D = 0.03 eV and a = 4.5 A−1. Our system of units (amu, A, eV)
defines a time unit (t.u.) equal to 1.018 × 10−14 s.
To describe the motions of the two strands, we introduce the new variables un and vn such that
un =
xn + yn√2
and vn =
xn − yn√2
,
(4)
where un and vn represent the in-phase and the out-of-phase motions. Taking into account Eq. (2) and Eq. (4), the
Hamiltonian of the system can be rewritten as
H =
N
Xn n 1
2
m u2
n +
1
2Xl=1
Jl(un − un−l)2o +
N
Xn n 1
2
m v2
n +
1
2Xl=1
Jl(vn − vn−l)2 + D(cid:16)e−a√2vn − 1(cid:17)2o.
The equations of motions of the system then read
mun =Xl=1
Jl (un+l − 2un + un−l) ,
mvn =Xl=1
Jl (vn+l − 2vn + vn−l) + 2√2aDe−a√2vn(cid:16)e−a√2vn − 1(cid:17) .
(5)
(6)
(7)
For a more realistic study of dynamical properties of DNA, one must take into account its environment. In the present
work, we take into account the Stokes (F st) and the long-range hydrodynamical (F hy) damping forces in the equations
of motion of the system. These forces account respectively for DNA molecules in a viscous environment and their inner
mechanism. The Stokes damping forces is given by
n = −mγst qn,
F st
(8)
where γst is the Stokes damping constant. The coordinate qn can be replaced by un or vn. In previous works in discrete
lattices, investigators assume the hydrodynamical damping force in the nearest neighbor interactions [25, 26, 27, 28]. In
this work, the DNA molecule is considered as a collection of nucleotides linked to the neighbors of the same strand by
spring. Each of them is assumed to be point masses of mass m. Thus, the displacement of one base pair causes a more or
less significant displacement of the other base pairs of the chain according to whether they are closed or distanced from
the initial base pair. This displacement gives rise to the hydrodynamical viscous forces which influence the motion of the
3
initial nucleotide. Since our work focuses on LRI between base pairs, we have introduced the LRI in the hydrodynamic
dissipation F hy in order to takes into account the contribution of all base pairs of the chain so that,
F hy
n = mXl=1
γl ( qn−l − 2 qn + qn+l) ,
(9)
where γl = γhyl−s′
defined above, we obtain the following equations of motion:
and γhy is the hydrodynamical damping coupling constant. Taking into account F st
n and F hy
n as
Jl
m
un =Xl=1
(un+l − 2un + un−l) − γSt un +Xl=1
γl ( un+l − 2 un + un−l)
Jl
m
(vn+l − 2vn + vn−l) +
2√2aD
m
vn =Xl=1
e−a√2vn (e−a√2vn − 1) − γSt vn +Xl=1
γl ( vn+l − 2 vn + vn−l) .
(10)
(11)
The solution un(t) of Eq. (10) is an ordinary solution of a damped linear schrodinger equation and represents a plane
wave in a viscous medium in the presence of LRI. In what follows, the system will be considered heavily damped. Our
investigations will be limited to the analysis of the dynamical behavior of the stretching motion of each base pair represented
by the solution of Eq. (11), in the presence of LRI and the "big viscosity " [10, 11].
3. Discrete difference operator technique
In this section, the DDO technique which is appropriate for long-range interacting systems [20] is used to study the
dynamics of DNA breathing. Assuming as usual small amplitude oscillation of the nucleotide around the bottom of the
Morse potential, we obtain up to the third order of the Morse potential the following equation of motion:
Jl
m
vn =Xl=1
m , α = − 3a
√2
(vn+l − 2vn + vn−l) − ω2
g(vn + αv2
n + βv3
γl ( vn+l − 2 vn + vn−l) .
(12)
n) − γSt vn +Xl=1
where ω2
g = 4a2D
3 . Eq. (12) describes the dynamics of the out-of-phase motion of the DNA in
viscous medium in the presence of LRI forces. Introducing the distance of neighboring bases r and assuming plane wave
and β = 7a2
solutions of the form
vn = F1ei(qnr−ωt) + c.c.
(13)
and substituting them into the equations of motion, we obtain the nonlinear dispersion relation in rotating wave approxi-
mation for the normal mode frequencies ωn and wave numbers qn
n = ω2
ω2
g(1 + 3βF12) + 4Xl=1
Jl
m
sin2(q0
nlr/2) − iωnγn,
where γn is the damping coefficient given by:
γn = γst + 4Xl=1
γl sin2(q0
nrl/2).
(14)
(15)
After some algebras, this dispersion relation can be rewritten as the sum of its real part ωr and imaginary part ωi. That
is
ωn = ωr + iωi,
ωr = ω0,np1 − δ2
n,
4
ωi = −
γn
2
,
(16)
with δn = γn
2ω0,n
and ω0,n the optical frequency of vibrations of base pairs in the absence of damping forces given by
ω2
0,n = ω2
g(1 + 3βF12) + 4Xl=1
Jl
m
sin2(q0
nrl/2).
(17)
We plot in Figure 1 the real and imaginary parts of the angular frequency of the wave (Eq. (16)), the real and imaginary
parts of the dispersion coefficient in the linear limit F1 → 0 for discretized values of the wave vector q0
γ0 = 0.15. In the panel (a) we observe that the real part of the angular frequency is equal to zero for qnr ∈] π
different from zero otherwise, namely qnr ∈ [0, π
the DNA molecule only if the carrier wave vector qnr is selected in a finite interval n[0, π
12 [ and
12 , 2π]. Then the vibration can appears and propagates in
The contribution of the long range decays of the stacking and viscous interactions are not the same, since the origins of
12 ] and qnr ∈ [ 23π
12 , 2π]o.
nr for s = 3.00 and
12 ] ∪ [ 23π
12 , 23π
the two forces are physically different, but nevertheless for seek of simplicity the same exponent is assumed that is s = s′.
Also, as in [29], we set γ0 = γst = γhy. At small wavelength, the soliton solution of Eq. (12) is found as an expansion in
normal modes and may be found in the form [30].
vn = ε[F1(z1, τ )ei(q0 rn−ω0t) + c.c.],
with
F1(z1, τ ) =
N
Xn=1
Bnei(δqn z1−δωnτ ),
qn = q0 + εδqn,
ωn = ω0 + ε2δωn.
(18)
(19)
The function F1 is a slowly varying function in space z1 = εrn and time τ = ε2t. The parameter q0 is the wavenumber of
the wave packet and the associate frequency ω0 ≡ ω(q0, F1 = 0) in the limit F1 → 0
The time derivative of the wave amplitude Eq. (19) reads:
∂F1(z1, τ )
∂τ
= [−iδωn]
N
Xn=1
Bnei(δqnz1−δωnτ ) = [−iδωn]F1.
(20)
From Eqs. (14), (19) and (18) it is seen that the term ε2δωn is an evolution function of two variables: the wavenumber q0
and the slowly varying wave amplitude εF12. The Taylor expansion of this term around the value q0 and εF1 = 0 and
neglecting higher order terms (> 2), give us
ε2δωn(∂qn,εF12) = δω0(q0) + (qn − q0)
∂ω0(q0)
∂q0 +
1
2
(qn − q0)2 ∂2ω0(q0)
(∂q0)2 + εF12 ∂ωn(q0)
.
(21)
The first term of the right hand site of Eq. (21) is assuming to be very close to zero. From Eq. (19) we have εδqn = qn− q0.
The above considerations in Eq. (21) lead to:
∂(F12)(cid:12)(cid:12)(cid:12)F1=0
δωn(∂qn,F12) =
(εδqn)
ε2
∂ω0(q0)
∂q0 +
1
2
(εδqn)2
ε2
∂2ω0(q0)
∂q02 + F12 ∂ωn(q0)
.
(22)
∂(F12)(cid:12)(cid:12)(cid:12)F1=0
The discrete difference operator is used instead of the continuous derivatives which can cause divergences. Therefore we
have:
∂ω0(q0)
∂q0 =
ω0(q0 + h) − ω0(q0)
h
,
∂2ω0(q0)
∂q02 =
ω0(q0 + h) − 2ω0(q0) + ω0(q0 − h)
h2
,
(23)
and finally we get,
δωn(∂qn,F12) =
(εδqn)ν
∆(ν)
h [ω0(q0)]
ε2ν!
hν
2
Xν=1
+ F12 ∂ωn(q0)
∂(F12)(cid:12)(cid:12)(cid:12)F1=0
,
5
(24)
where ∆(ν)
h
is the difference operator of order ν with step size h = 2π/N in the limit F1 = 0 and given below,
∆(1)
h [ω0] = ω0(q0 + h) − ω0(q0),
∆(2)
h [ω0] = ω0(q0 + h) − 2ω0(q0) + ω0(q0 − h).
From Eq. (19) the term (δqn)ν can be expressed as follows:
(iδqn)ν F1 =
∂νF1
∂zν
1
.
Using Eqs. [26-23] into Eq. (20) give the nonlinear equation of evolution of the envelope function written as
where the parameters vg, P and Q are the group velocity, the dispersion and the nonlinearity coefficients given by
ih ∂F1
∂τ
+
vg
ε
∂F1
∂z1i + P
∂2F1
∂z2 + QF12F1 = 0
vg =
∆(1)
h [ω0]
h
,
P =
∆(2)
h [ω0]
2h2
,
The above parameters can be rewritten as:
vg = vgr + ivgi,
vgr =
Q = −
∂ωn(q0)
∂(F12)(cid:12)(cid:12)(cid:12)F1=0
.
∆(1)
h [ω0
i ]
h
∆(2)
h [ω0
i ]
2h2
,
,
vgi =
Pi =
Qi = 0.
∆(2)
∆(1)
h [ω0
r ]
h
h [ω0
r ]
2h2
3βω2
g
2ω0
r
,
,
P = Pr + iPi,
Pr =
Q = Qr + iQi,
Qr = −
(25)
(26)
(27)
(28)
(29)
(30)
Setting ξ1 = z1 − εvgt in the co-moving reference frame with a rescaled time t such as t → ε2t, Eq. (24) becomes
i
∂F1
∂t
+ (Pr + iPi)
∂2F1
∂ξ2
1
+ QF12F1 = 0.
It should be noted that Eq. (30) is the well-known CGL equation for the evolution of the envelope where the dispersion
coefficient is complex and the nonlinearity coefficient is real. Similar equation was found in Ref. [10, 11] where the authors
studied the dynamics of a damped DNA in the absence of hydrodynamical damping and LRI forces using the semi-discrete
approximation. In their study they found the dispersion coefficient real and the nonlinearity coefficient complex contrary
of the one obtained in this work. The nonlinearity coefficient Q and the dispersion coefficient P not only depend on
the wave vector q0
nr, and the Stokes viscous forces as previously mentioned by these authors, but also depend on the
hydrodynamic damping and the LRI forces.
Several methods related to soliton solutions for the specific forms of CGL equation have been developed [21, 31, 32].
A key problem in this paper is to give an analytical soliton solution of the CGL equation (Eq. (30)), and use it to study
the effect of viscosity and LRI on the DNA opening state configuration. The character of this solution is determined by
the sign of Q and Pr while the stability of the plane wave solution through the Benjamin-Feir instability depends on the
sign of the product PrQ. For PrQ < 0, the plane wave solution is stable and for PrQ > 0 it is unstable. Particularly,
since the nonlinear coefficient Q is always negative the sign of the constant PrQ depends on real part of the dispersion
coefficient Pr which can take positive or negative values depending on the range of variations of the wave vector. Here,
only localized solutions in space for any wave carrier whose wavenumber is in the positive range of PrQ are considered.
6
In Figure 2, the product PrQ is represented as a function of the wave vector for s = 3.00 and γ = 0.15. We observe
in the plots that PrQ > 0 is always positive. From Eqs. (15), (16) and (29), we observe that the imaginary parts of the
solitonic parameters strongly depend on the damping forces of the system, therefore their absolute values decrease with
the decreasing of the damping constant and vanish when the damping is switched off.
As in [21, 31, 32], an analytical solution of Eq. (31), is found in the form
F1 = Ah sech(ηξ1)i1+iσ
e−iφt,
(31)
where A, φ, η−1, and σ are parameters to be determined and represent respectively the complex amplitude, the complex
"angular frequency ", the width and the chirp of the soliton. By introducing Eq. (31) into Eq (30) and after canceling the
terms in [ sech(ηξ1)i1+iσ
, the real and imaginary part of the phase of the soliton is written
φr = −η2h(1 − σ2)Pr − 2σPii,
φi = −η2h(1 − σ2)Pi + 2σPri.
, the width and the chirp of the soliton is given by
From the annihilation of the terms in [ sech(ηξ1)i3+iσ
where ∆ = 9P 2
r + 8P 2
i and σ a solution of the following quadratic equation
A = ηs(cid:12)(cid:12)(cid:12)
(2 − σ2)Pr − 3σPi
Qr
,
σ =
3Pr + √∆
2Pi
(cid:12)(cid:12)(cid:12)
Piσ2 − 3Prσ − 2Pi = 0.
(32)
(33)
(34)
(35)
(36)
(37)
This choice of σ implies that the soliton is strongly chirped.
Now to determine the complex amplitude Aγ of the soliton, let us consider the system in the non-viscous limit (γ0 = 0).
In that case Pi vanishes and Eq. (30) becomes the standard NLS equation
where the associated group velocity, dispersion coefficient and nonlinearity coefficient are
i
∂G1
∂t
+ P ′
∂2G1
∂ξ2
1
+ Q′G12G1 = 0,
=
∆(1)
h [ω0
0]
h
,
v′g ≡ vgr(cid:12)(cid:12)(cid:12)γ=0
=
∆(2)
h [ω0
0]
2h2
,
P ′ ≡ Pr(cid:12)(cid:12)(cid:12)γ=0
3βω2
g
2ω0
.
= −
Q′ ≡ Qr(cid:12)(cid:12)(cid:12)γ=0
The solution of Eq. (35) is the well known modulated solitonic wave called breather [24, 33] and given by
G1 = A′ sechhL(ξ1 − uet)iei ue
2P ′ (ξ1−uct),
where ue and uc are real parameters representing respectively the velocities of the envelope and the carrier wave of the
soliton. The amplitude of the envelope A′ and its inverse width L′ are given by the relations
L′ = pu2
e − 2ueuc
2P ′
,
A′ =s u2
e − 2ueuc
2P ′Q′
.
(38)
Let us assume that at the initial time (t = 0), for γ0 = 0, the soliton solution Eqs. (31) and (38) should be equivalent,
A(cid:12)(cid:12)(cid:12)γ=0
= A′ei ue
2P ′ ξ1,0 ,
7
(39)
where ξ1,0 is the initial position of the soliton. Taking into account this new expression of the amplitude A, the one soliton
solution of Eq. (30) is
F1 = ηs(cid:12)(cid:12)(cid:12)
(2 − σ2)Pr − 3σPi
Qr
eφith sech(ηξ1)i1+iσ
(cid:12)(cid:12)(cid:12)
ei( ue
2P ξ1,0−φr t).
(40)
Now considering the fact that the angular frequency of the soliton is complex (see Eq. (16)), one can note that the term
eωit enters inside the amplitude of the soliton. Thus, the complete expression of the envelope soliton written in the original
temporal (t) and spacial (z) variables is finally given by
F ′1 =F1eωit = ηs(cid:12)(cid:12)(cid:12)
ξ1,0 + η2h(1 − σ2)Pr − 2σPiit,
with
Θ =
ue
2P
(2 − σ2)Pr − 3σPi
Qr
e−Γth sech(cid:2)η(z1 − εvgrt)(cid:3)i1+iσ
eiΘ
(cid:12)(cid:12)(cid:12)
Γ =
γn
2
+ η2h(1 − σ2)Pi + 2σPri,
where Γ is the effective damping constant of the medium.
e − 2ueuc
η = pu2
(cid:12)(cid:12)(cid:12)
(2 − σ2)Pr − 3σPi(cid:12)(cid:12)(cid:12)
To obtain the solution of the EOM of the out-of-phase motion vn given by Eq. (11), some results of the previous
section are used. Using Eqs. (16), (28) and (40) the soliton solution describing the out-of-phase motion of the DNA in
the viscous medium takes the form
where
vn(t) =2εAe−Γt sech(cid:2)εη(nr − vgrt)(cid:3) cos(qγnr − t)
nr (cid:17) + (σ/nr) log(cid:12)(cid:12)(cid:12)
2P(cid:16) ξ1,0
qγ = q0 +
ue
= ω0
r − η2h(1 − σ2)Pr − 2σPii,
sech(cid:2)εη(nr − vgrt)(cid:3)(cid:12)(cid:12)(cid:12)
.
The solution Eq. (43) represents the breather solution in the DNA molecule suggested by the Infrared and Raman
experiments [34]. This solution is represented in Figure 3. It can be seen that due to the damping effect, the amplitude
of the soliton is a decreasing function of time and hence it will propagates only for a limited distance and vanishes.
Prohofsky et al. have shown in their studies [34] that, the breather can be strongly located, or distributed on a wide
zone of the DNA molecule and could be at the origin of the localization of the energy in the molecule which lead to the
local denaturation.
It is noteworthy to mention that if one uses the semi-discrete approach [10, 11], the results fails in the derivation
of the breather soliton profile Eq. (43).
In fact, we can obtain a similar CGL equation Eq. (30), but with a different
dispersion coefficient containing continuous derivatives instead of difference operators as given in equation Eq. (28). Using
the semi-discrete approach, we obtain:
(41)
(42)
(43)
(44)
1
γl(rl)2 cos(qrl),
Pr =
1
r
, ω2 = ω2
+ ωiγl(cid:17)(rl)2 cos(qlr) − vgγ2i, Pi = −
2Xl=1
sin2(qrl/2), ωi = −
2ωrhXl=1(cid:16) Jl
ωr = ωr1 −(cid:16) γ
2ω(cid:17)2
ωr Xl=1(cid:16) Jl
+ ωiγl(cid:17)l sin(qrl), vgi = −rXl=1
(−2α +
β/α + BC), Qi = −
g + 4Xl=1
CD1 B = 4ω2
Qr = −
γll sin(qrl),
ω2
gα
ωr
ω2
gα
ωr
vgr =
Jl
m
3
2
m
m
γl sin2(qrl/2),
(45)
γ
2
, γ = γst + 4Xl=1
mXl=1
g +
ς = ω2
4
r ς, D1 = 2γωr, C =
Jl sin2(qrl),
ω2
gα
B2 + D2
1
.
8
In Eq. (45), the appearance of continuous derivatives can cause a divergence of both the group velocity vg and the dispersion
coefficient P . The coefficients oscillate for different values of chain length and does not converge to a definite value.
4. Numerical investigations.
The results discussed in the previous section are obtained from the CGL equation (Eq. (30)) derived after some
approximations and hypothesis and not from the discrete EOM. In order to verify the analytical predictions and check if
the above analytical breather soliton can survive in the discrete lattice, the numerical simulations of the discrete EOM
(Eq. (11)) is carried out by means of a standard fourth-order Runge-Kutta computational scheme with periodic boundary
conditions. In our simulations we use the initial condition given by Eq. (43) and time step h = 2π/N t.u. with N = 600.
Figure 4 presents the 3D and 2D time-evolution of the solution for a value of LRI parameter s = 3.00, and we observe
the decreasing of the amplitude of the soliton due to the damping forces (Figure 4a). Figure 4b presents the aspect of the
numerical solution at t = 115. As predicted by the analytical results, we observe that the breathing mode appears in DNA
molecule when the wave vector is small, namely qnr ≤ π
12 which corresponds to the domain where the real parts of the
angular frequency is different from zero. In Figure 5 and Figure 6, we have depicted the 2D schematic representation of
the analytical and numerical solutions at few time positions for two different wave vectors (qnr = π/16 and qnr = π/18):
t = 0, t = 50, t = 200 and t = 300. We observe that when the wave vector increases, the width of the soliton grows by
increasing the number of base pairs in the bubble. At the same time, the wave amplitude must increase in order to keep the
soliton within the lattice length limits. Therefore, small wavenumber leads to a more localized solution. The decreasing
of the amplitude of the soliton in time is observed. Also we notice that the shape, the decay of the amplitude and the
number of base pairs in the bubble of both solutions after a limited time propagation are the same. But after a long time
propagation the number of base pairs which form the bubble remains constant, the shape of the numerical solution is
slightly modified also the decay in the amplitude of the numerical solution is less than the theoretical expectations, due to
the discreteness effects which usually tend to slow down the motion [9]. These results confirm that our analytical solution
is stable and suitable to predict formation of breather solitons in the DNA molecular chain in the presence of damping
and LRI forces.
5. Conclusion
In this work, we have studied the dynamics of breather solitons in a long-range version of the Peyrard-Bishop DNA
model taking into Stokes and hydrodynamical viscous forces. Using the discrete difference operator technique, we have
shown that the out-of-phase motion can be described by the CGL equation. As compared to the semi-discrete approach,
this technique can lead to the correct expression for the soliton parameters. The breather soliton which represents the
opening of base pairs experimentally observed in the DNA molecular chain in the form of bubble, has been found stable
when it propagates, however its amplitude decreases due to damping effect. Our numerical simulations have confirmed
the validity of the analytical approximate results.
9
References
References
[1] S.W. Englander et al., Proc. Natl. Acad. Sci. U.S.A. 77, 7222 (1980).
[2] S. Yomosa, Phys. Rev. A 27, 2120 (1983).
[3] S. Homma and S. Takeno, Prog. Theor. Phys. 70, 308 (1983);.
[4] S. Takeno and S. Homma, Prog. Theor. Phys. 72, 679 (1984).
[5] M. Peyrard and A.R. Bishop, Phys. Rev. Lett. 62, 2755 (1989).
[6] T. Dauxois, M. Peyrard, and A.R. Bishop, Phys. Rev. E 47, R44 (1993).
[7] T. Dauxois and M. Peyrard, Phys. Rev. E 51, 4027 (1995).
[8] A. Campa and A. Giansanti, Phys. Rev. E 58, 3585 (1998).
[9] T. Dauxois, Phys. Lett. A. l59, 390 (l991).
[10] S. Zdravkovi´c and M.V. Satari´c, Chin. Phys. Lett. 24, 1210 (2003).
[11] S. Zdravkovi´c M.V. Satari´c and L. Hadzievski, Chaos 20, 043141 (2010).
[12] C. Calladine et al., Understanding DNA, 3rd edition (Academic Press, London, 2004).
[13] N.N. Shafranovskaya et al., Pis'ma. Zh. Eksp. Teor. Fiz. 15, 404 (1972).
[14] D.C. Rau and V.A. Parsegian, Biophysics J. 61, 246 (1992).
[15] Y. Ishimori, Prog. Theor. Phys. 68, 402 (1982).
[16] Yu. B. Gaididei, S.F. Mingaleev, P.L. Christiansen and K. ∅. Rasmussen, Phys. Lett. A. 222, 152 (1996).
[17] S. F. Mingaleev, Yu.B. Gaididei, and F.G. Mertens, Phys. Rev. E. 58, 3833 (1998).
[18] A. Mvogo, G.H. Ben-Bolie and T.C. Kofan´e, Phys. Lett. A. 378, 2509 (2014).
[19] A. Mvogo, G.H. Ben-Bolie and T.C. Kofan´e, Chaos 25, 063115 (2015).
[20] G. Miloshevich, J.P. Nguenang, T. Dauxois, R. Khomeriki and S. Ruffo, J. Phys. A: Math. Theor. 50, 12LT02
(2017).
[21] N.R. Pereira and L. Stenflo, Phys. Fluids 20, 1733 (1977).
[22] S. Flach, Phys. Rev. E 58, R4116 (1998).
[23] Y.B. Gaididei, S.F. Mingaleev, P.L. Christiansen and K.∅ Rasmussen, Phys. Rev. E 55, 6141 (1997).
[24] M. Peyrard, Nonlinearity 17, R1 (2004).
10
[25] E. Ar´evalo and F.G. Mertens, Phys. Rev. E. 67, 016610 (2003).
[26] C. Brunhuber and F.G. Mertens, Phys. Rev. E. 73, 016614 (2006).
[27] I. Daumont and M. Peyrard, Chaos 13, 624 (2003).
[28] M. Peyrard and I. Daumont, Europhys. Lett., 59 834 (2002).
[29] E. Ar´evalo Yu. Gaididei and F.G. Mertens, Eur. Phys. J. B 27, 63 (2002)
[30] J.W. Boyle, S. A. Nikitov, A.D. Boardman, J.G. Booth and K. Booth, Phys. Rev. B 53, 12173 (1996)
[31] L.M. Hocking and K. Stewartson, Proc. R. Soc. London, Ser. A 326, 289 (1972).
[32] N. Akhmediev and A. Ankiewicz: Solitons of the Complex Ginzburg-Landau Equation in Spatial Solitons,
Edited by S. Trillo, p. 311 (Springer, New York, 2002).
[33] A.C. Scott, F.Y.F. Chu and D.W. McLaughlin, Proc. IEEE. E. 61, 1443 (1973).
[34] B.F. Putnam, L.L. Van Zandt, E.W. Prohofsky, K.C. Lu, and W. N. Mei, Biophys. J. 35, 271 (1981); E.W.
Prohofsky, K.C. Lu, L.L. Van Zandt and B.F. Putnam, Phys. Lett. A 70, 492 (1979).
0.05
0.045
0.04
0.035
0.03
0.025
0.02
0.015
0.01
0.005
)
1
-
.
u
.
t
(
r
ω
0
0
π/4
π/2
3π/4
(a)
π
q
r
n
(c)
)
.
u
.
t
/
2
Å
(
r
P
0
-2
-4
-6
-8
-10
-12
-14
-16
5π/4
3π/2
7π/4
2π
(b)
π
q
r
n
(d)
π/4
π/2
3π/4
-0.05
-0.1
-0.15
-0.2
)
1
-
.
u
.
t
(
i
ω
-0.25
-0.3
-0.35
-0.4
0
0.1
0
-0.1
-0.2
-0.3
-0.4
)
.
u
.
t
/
2
Å
(
i
P
5π/4
3π/2
7π/4
2π
0
π/4
π/2
3π/4
π
q
r
n
5π/4
3π/2
7π/4
2π
-0.5
0
π/4
π/2
3π/4
5π/4
3π/2
7π/4
2π
π
q
r
n
Figure 1: (Color online) (a) and (b) the real and imaginary part of the dispersion relation (Eq. (16)) in the linear limit F1 → 0. (c) and (d) the
real and imaginary part of the second discrete derivative of the dispersion relation (Eq. (29)) (Dispersion coefficient) in terms of the discretized
values of the wave vector qnr = 2πn/N , N = 600 for s = 3.00, γ0 = 0.15 t.u−1.
11
Q
P
r
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
π/4
π/2
3π/4
π
q
r
n
5π/4
3π/2
7π/4
Figure 2: (Color online) The product P rQ in terms of the the discretized values of the wave vector qnr = 2πn/N , N = 600, for s = 3.00,
γ0 = 0.15 t.u−1
)
Å
(
n
v
0.08
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
0
(b): t=115
100
200
300
400
500
600
Base pairs (n)
Figure 3: (Color online) (a) Analytical stretching of the nucleotide pair as a function of the time and the number of base pairs. (b) stretching
of the nucleotide pair as a function of the number of base pairs at t = 115 for s = 3.00, ε = 0.9, ue = 1, γ0 = 0.15 t.u−1, uc = 0.45ue and
q0
nr = π
16 .
12
)
Å
(
n
v
0.08
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0
(b): t=115
100
200
300
400
500
600
Base pairs (n)
Figure 4: (Color online) (a) Numerical stretching of the nucleotide pair as a function of the time and the number of base pairs. (b) stretching
of the nucleotide pair as a function of the number of base pairs at t = 115 for s = 3.00, ε = 0.9, ue = 1, γ0 = 0.15 t.u−1, uc = 0.45ue and
q0
nr = π
16 .
)
Å
(
n
v
)
Å
(
n
v
0.1
0.08
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0
0.08
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0
(a): t=0
100
200
300
400
500
600
Base pairs (n)
(c): t=200
100
200
300
400
500
600
Base pairs (n)
)
Å
(
n
v
)
Å
(
n
v
0.1
0.08
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0
0.08
0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0
(b): t=50
100
200
300
400
500
600
Base pairs (n)
(d): t=300
100
200
300
400
500
600
Base pairs (n)
Figure 5: (Color online) Comparison between analytical and numerical solution of the EOM Eq. (11) at different time positions for s = 3.00,
ε = 0.9, ue = 1, γ0 = 0.15 t.u−1, uc = 0.45ue and q0
nr = π
16 . (solid blue line) the numerical solution. (dash red line) the analytical solution.
13
)
Å
(
n
v
0.12
0.1
0.08
0.06
0.04
0.02
0
-0.02
-0.04
0
0.1
)
Å
(
n
v
0.05
0
-0.05
0
(a): t=0
100
200
300
400
500
600
Base pairs (n)
(c): t=200
100
200
300
400
500
600
Base pairs (n)
)
Å
(
n
v
0.12
0.1
0.08
0.06
0.04
0.02
0
-0.02
-0.04
0
0.1
)
Å
(
n
v
0.05
0
-0.05
0
(b): t=50
100
200
300
400
500
600
Base pairs (n)
(d): t=300
100
200
300
400
500
600
Base pairs (n)
Figure 6: (Color online) Comparison between analytical and numerical solution of the EOM Eq. (11) at different time positions for s = 3.00,
ε = 0.9, ue = 1, γ0 = 0.15 t.u−1, uc = 0.45ue and q0
nr = π
18 . (solid blue line) the numerical solution. (dash red line) the analytical solution.
14
|
1709.02703 | 3 | 1709 | 2018-01-30T10:33:27 | Hydro-osmotic instabilities in active membrane tubes | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.SC"
] | We study a membrane tube with unidirectional ion pumps driving an osmotic pressure difference. A pressure driven peristaltic instability is identified, qualitatively distinct from similar tension-driven Rayleigh type instabilities on membrane tubes. We discuss how this instability could be related to the function and biogenesis of membrane bound organelles, in particular the contractile vacuole complex. The unusually long natural wavelength of this instability is in agreement with that observed in cells. | physics.bio-ph | physics |
Hydro-osmotic instabilities in active membrane tubes
Sami C. Al-Izzi,1, 2, 3, 4 George Rowlands,2 Pierre Sens,3, 4 and Matthew S. Turner2, 5
1Department of Mathematics, University of Warwick, Coventry CV4 7AL, UK
2Department of Physics, University of Warwick, Coventry CV4 7AL, UK
3Institut Curie, PSL Research University, CNRS,
Physical Chemistry Curie, F-75005, Paris, France
4Sorbonne Universit´es, UPMC Univ Paris 06, CNRS, UMR 168, F-75005, Paris, France
5Centre for Complexity Science, University of Warwick, Coventry CV4 7AL, UK
We study a membrane tube with unidirectional ion pumps driving an osmotic pressure difference.
A pressure driven peristaltic instability is identified, qualitatively distinct from similar tension-
driven Rayleigh type instabilities on membrane tubes. We discuss how this instability could be
related to the function and biogenesis of membrane bound organelles, in particular the contractile
vacuole complex. The unusually long natural wavelength of this instability is in agreement with
that observed in cells.
The "blueprint" for internal structures in living cells
is genetically encoded but their spatio-temporal organi-
sation ultimately rely on physical mechanisms.
A key contemporary challenge in cellular biophysics
is to understand the physical self-organization and reg-
ulation of organelles [1, 2]. Eukaryotic organelles bound
by lipid membranes perform a variety of mechanical and
chemical functions inside the cell, and range in size, con-
struction, and complexity [3]. A quantitative under-
standing of how such membrane bound organelles func-
tion have applications in bioengineering, synthetic biol-
ogy and medicine. Most models of the shape regulation
of membrane bound organelles invoke local driving forces,
e.g. membrane proteins that alter the morphology (of-
ten curvature) [4 -- 6]. However other mechanisms, such as
osmotic pressure, could play an important role [7].
Membrane tubes are ubiquitous in cells, being found in
organelles such as the Golgi and endoplasmic reticulum
[3] and elsewhere. Models for their formation typically
involve the spontaneous curvature of membrane proteins
[5] or forces arising from molecular motors, attached to
the membrane, that pull tubular tethers as they move
along microtubules [8]. Many of these tubules may con-
tain trans-membrane proteins that can alter the osmotic
pressure by active transport of ions. Most work on the
biogenesis of cellular organelles has focused on their static
morphology and generally not on their non-equilibrium
dynamics.
In what follows we consider an example in
which the out-of-equilibrium dynamics drives the mor-
phology, Fig. 1. Our study is inspired by the biophysics
of an organelle called the Contractile Vacuole Complex
but additionally reveals a new class of instabilities not
previously studied that are of broad, perhaps even uni-
versal, physiological relevance.
The Contractile Vacuole Complex (CVC) is an or-
ganelle found in most freshwater protists and algae that
regulates osmotic pressure by expelling excess water [10 --
14].
Its primary features is a main vesicle (CV) that
is inflated by osmosis and periodically expels its contents
through the opening of a large pore - probably in response
to membrane tension - connecting it to the extracellular
FIG. 1.
(a) Diagram of the contractile vacuole complex. The
tube is shown connected to the main body of the CV (left).
As ions are pumped in, increasing the osmotic pressure, the
tube undergoes a swelling instability and undulations develop
with some wavelength λ. This phenomena is observed in the
contractile vacuoles of, e.g. paramecium multimicronucleatum
[9, 10]. (b) Schematic of a membrane tube with ion pumps
and surface undulations. A cartoon of a representative ion
pump is shown in the top right.
environment, thereby regulating cell volume [9, 14]. Wa-
ter influx into the CVC is due to an osmotic gradient gen-
erated by ATP-hydrolysing proton pumps in the mem-
brane that move protons into the CVC [12, 15 -- 17]. In
many organisms such as Paramecium multimicronuclea-
tum, the CVC includes several membrane tubular arms
connected to the main vesicles, which are thought to be
associated with the primary sites of proton pumping and
water influx activity [18]. The tubular arms do not swell
homogeneously in response to water influx, but rather
show large undulatory bulges with a size comparable to
the size of the main CV, leading us to speculate that this
might even play a role in CV formation de novo. These
tubular arms appear to be undergoing a process simi-
lar to the Pearling or Rayleigh instability of a membrane
tube under high tension [19 -- 26] or an axon under osmotic
shock [27], but with a much longer natural wavelength:
Rayleigh instabilities have a natural wave length λ ∼ R
Ru(z)tPore(b)(a)Plasma membraneTubular armwhere R is the tube radius. Here we derive the dynamical
evolution of a membrane tube driven out-of-equilibrium
by osmotic pumping.
In the CVC, the tubular arms are surrounded by a
membrane structure resembling a bicontinuous phase
made up of a labyrinth tubular network called the smooth
spongiome (SS). We assume this to represent a reservoir
of membrane keeping membrane tension constant and
uniform during tube inflation.
It is possible to imple-
ment more realistic area-tension relations [26], however
this is beyond the scope of the present work.
The CVC is comprised of a phospholipid bilayer mem-
brane. This bilayer behaves in an elastic manner [28, 29].
At physiological temperatures these lipids are in the fluid
phase [3, 29]. For simplicity we will treat the bilayer as
a purely elastic, fluid membrane in the constant tension
regime, neglecting the separate dynamics of each leaflet.
The membrane free energy involves the mean curvature
H and surface tension γ [28, 30, 31]
(2H)2 + γ
∆P dV ,
(1)
(cid:90)
F =
dA
S
(cid:16) κ
2
where dA and dV are the area and volume elements on
S, κ is the bending rigidity, and ∆P is the pressure dif-
ference between the fluid inside and outside the tube.
Assuming radial symmetry and integrating over the vol-
ume of the tube we obtain
(cid:90) ∞
−∞
κ
2
F = 2π
dz
r
1(cid:113)
(cid:113)
+γr
1 + (∂zr)2
∂zzr
1 + (∂zr)2 − 1
(cid:21)
r
1 + (∂zr)2 − 1
2
r2∆P
(cid:90)
(cid:17) −
(cid:32)
where r(z, t) is the radial distance of the axisymmetric
membrane from the cylindrical symmetry axis and z mea-
sures the coordinate along that axis, see S.I. for details.
We use Eq. (2) as a model for the free energy of a radial
arm of the CVC. Ion pumps create an osmotic pressure
difference that drive a flux of water to permeate through
the membrane. We calculate the dominant mode of the
hydro-osmotic instability resulting from the volume in-
crease of the tube lumen. We write the radius of the tube
as r(z, t) = R + u(z, t), with u assumed small, and make
q ¯uqeıqz.
Absorbing the q = 0 mode into R = R(t) allows us to
use of the Fourier representation u(z, t) = (cid:80)
write(cid:82) u dz = 0. The free-energy per unit length can be
written at leading order as
F = F (0) +
α(q)¯uq2
(3)
π
R
(cid:88)
q
(cid:19)
+ 1
2
(cid:18) κ
2R
(cid:18)
where
α(q) =
κ
R2
and
(qR)4 − (qR)2
+ γ(qR)2 − ∆P R (4)
F (0) = 2π
+ γR − 1
2
∆P R2
(5)
(cid:19)
2
√
(cid:113) κ
Identifying the static pressure difference ∆P with the
Laplace pressure PL = −κ/(2R3) + γ/R, the point at
which the q = 0 mode goes unstable can be identified:
3Req
the membrane tube is unstable for tube radii R >
where Req =
2γ is the equilibrium radius of a tube with
∆P = 0. This criterion for the onset of the instability is
the same as the Rayleigh instability on a membrane tube
[24], however the instability is now driven by pressure
not surface tension. This is a crucial difference. It leads
to a qualitatively different evolution of the instability,
as we now show.
In what follows we are interested in
the dynamics of the growth of unstable modes after the
3Req. Our initial condition
cylinder has reached radius
is a tube under zero net pressure, although the choice
of initial condition is not crucial. We assume that the
number of proton pumps moving ions from the cytosol
into the tubular arm depends only on the initial surface
area, i.e. it is fixed as the tube volume (and surface)
varies.
√
We denote the number of ions per unit length in the
tube as n and write an equation for the growth of n as
dn
dt
=
0,
2πβReq,
t ∈ (−∞, 0)
t ∈ [0,∞)
(6)
(cid:40)
(cid:33)2
where β is a constant equal to the pumping rate of a sin-
gle pump multiplied by the initial area density of pumps.
The density of ions, ρI , can be obtained by solving
Eq. (6) and dividing by volume per unit length, v(t),
(2)
ρI =
n(t)
v(t)
=
n0
v(t)
+
2πβReqt
v(t)
.
(7)
The growth of the tube radius is driven by a differ-
ence between osmotic and Laplace pressure [32]. This
means the rate equation for the increase in volume can be
written in terms of the membrane permeability to water.
Assuming that the water permeability (number of wa-
ter permeable pores) is constant during tube inflation,
we write the volume permeability per unit tube length
µ(cid:48) = 2πReqµ, where µ is the (initial) permeability of the
membrane. Thus
= µ(cid:48) (kBT (ρI − ρI (t = 0)) − P )
(8)
dv
dt
where the osmotic pressure is approximated by an ideal
gas law. This can be transformed into an equation for
R(t) on the time interval t ∈ [0,∞). We identify P with
the Laplace pressure. This leads to
(cid:18)
(cid:18) t
R2
(cid:19)(cid:18) 1
(cid:19)(cid:19)
d R
dt
=
τpump
τµ
1
R
+
1 +
γ
R
− 1
R2
(9)
γ
kB T ReqρI (t=0) , τpump = ReqρI (t=0)
where γ =
,
τpump
R = R
µ(cid:48)kB T ρI (t=0) . τpump and τµ represent
Req
the time-scales of pumping and permeation of water re-
spectively. The experimental time-scale for radial arm
and τµ =
, t = t
Req
2β
inflation is consistent with a value of τpump ∼ 1 − 10−1s.
These dynamics assume our ensemble conserves surface
tension, not volume (as in the usual Rayleigh instability).
This proves to be a crucial difference.
Values of Req = 25nm, γ = 10−4N m−1 and hence
κ are estimated using experimentally measured values
from [33, 34]. We take a typical ionic concentration in
the cytosol of a protist for ρI (t = 0) = 3.0 × 108µm−3
(around 10 mMol) [12, 29, 35]. Making an order of mag-
nitude estimate of β from the literature on the CVC
[12, 36, 37] leads to estimates of β ∼ 106-109µm−2s−1.
Temperature is taken as T = 310K. The permeabil-
ity of polyunstaurated lipid membranes is thought to be
around µ = 10−4µm Pa−1s−1 [38]. This permeability
could be much larger in the presence of water channels
but we find that our results are rather insensitive to in-
creasing the value of µ because, for physiological param-
eter values, our model remains in the rapid permeation
regime, i.e. τµ/τpump (cid:28) 1. This permits a multiple time-
scales expansion [39] of Eq. (9). With γ ∼ 10−3 (cid:28) 1 we
find the approximate asymptotic solution
(cid:19)1/2
(cid:18) τµ
(cid:19)
τpump
+ O
.
(10)
(cid:18) t
R(t) =
+ 1
τpump
This solution agrees well with numerical solutions to
Eq. (9). Using Eq. (10) and Eq. (4) we can compute
the time at which each q mode goes unstable, see S.I.
We now proceed to deriving the dynamical equations
for the Fourier modes. The equations governing the sol-
vent flow are just the standard inertia free fluid equations
for velocity field (cid:126)v. These are the continuity and Stokes
equations for incompressible flow
(cid:126)∇ · (cid:126)v = 0;
(cid:126)∇P = η∇2(cid:126)v
(11)
where P is the hydrodynamic pressure and η = 10−3Pa.s
the viscosity. The linearised boundary conditions are:
vrr=R = u + vp, where vp is the permeation velocity
(proportional to the hydrodynamic pressure jump across
the membrane: vp = µ∆Pr=R), and vzr=R = 0. The
second condition is justified by invoking the membrane
reservoir as a mechanism for area exchange.
Solving these equations and substituting into the mem-
brane force balance equation gives (in the small qR limit)
(cid:18) q2R(t)
8η
(cid:19)
¯uq = −αL(q)
+
2µR(t)
R3
eq
¯uq
(12)
where ¯uq is the Fourier representation of u in the z di-
rection (see S.I. for details). The response function αL
is obtained by replacing the static pressure difference by
the Laplace pressure PL in Eq. (4). Note that the term
involving µ, capturing mode growth due to permeation,
is only relevant for wavelengths λ > 100Req, hence we
will discard it in our analysis for simplicity (but retain
it in the numerics, for completeness). The growth rate
for a given mode is now time dependent, hence the mode
3
amplitude cannot be obtained from the maximum of the
growth rate, but depends on the growth history and must
be obtained by solving the full, time-dependent problem.
We identify the instability as being fully developed when
our linearised theory breaks down. We defined the dom-
inant mode of the instability, called q, as the first mode
with an amplitude reaching (cid:112)(cid:104)¯uq2(cid:105) = Req (a choice
that does not influence our results, see S.I.). This occurs
at t = tfinal.
The fluctuations of modes with wavenumber q about
the radius R(t) follow the dynamics of the Langevin equa-
tion based on Eq. (12)
η(q) ¯uq = −αL(q)¯uq + ζq
(13)
where η(q) = 8η
following statistical properties
Rq2 and ζq, the thermal noise, has the
(cid:104)ζq(cid:105) = 0
(cid:104)ζq(t1)ζq(cid:48)(t2)(cid:105) = δqq(cid:48)δ (t1 − t2)
(14)
(15)
kBT R
πη(q)
.
Here the thermal noise is found using the equipartition
theorem, and is thus integrated around the tube radius.
Solving this Langevin equation for (cid:104)¯uq2(cid:105), using an
initial condition of an equilibrium tube and the approxi-
mate form of R(t) (Eq. (10)) we find an integral equation
for the mode growth
(cid:104)¯uq2(cid:105)
R2
eq
=
2κπ(1 + q4)
kBT
(cid:90) t
e(F (0)−F (t))
+ e−F (t)
kBT q2(t(cid:48) + 1)
τpump
0
κπ
τη
eF (t(cid:48))dt(cid:48)
(16)
where t(cid:48) is a time variable integrating over the noise ker-
nel (in units of τpump), τη = 8R3
eqη/κ, q = qReq and
F (t) =
×(cid:16)
2τpump q2 R(t)
40 − 5t + q2 R(t)2(cid:16)
15τη
3t − 2 + 6q2 R(t)2(cid:17)(cid:17)
(17)
Integrating this numerically, together with Eq. 10, we
can find the dynamics of the modes. The distribution
of mode amplitude against q is shown in Fig. 2. Al-
though the smallest q modes go unstable first, they have
very slow growth and so the mode that dominates the
instability arises from the balance between going unsta-
ble early (favouring low q) and growing fast (favouring
higher values of q).
(cid:112)(cid:104)¯uq2(cid:105) = Req, see Fig. 3. This gives a dominant
We can compute numerically the natural wavelength
associated with the dominant mode, q, the first to reach
wavelength λ ∼ 100 Req ∼ 2µm for parameters consis-
tent with the CVC, much larger than that found in the
Rayleigh instability, but consistent with observations of
the CVC [10]. Understanding why this is the case is
not straightforward by inspection of the growth equation
4
FIG. 3. Plot of dominant wavenumber q = q∗Req of the insta-
bility against ratio of viscous to pumping timescales τη/τpump
and ratio of viscous to permeable timescales, τη/τµ. All other
parameters are the same as in Fig. 2. The blue rectangle
indicates typical physiological parameters.
ity that we analyse here recognises the presence of active
pumps in the organelle membrane, which can drive os-
motic changes in the organelle lumen [10]. There is some
correspondence between the fast pumping limit in Fig. 3
(τη/τpump large) and the osmotic shock situation. The
instantaneous growth rates have the same dependence in
the tube radius, but have a different time dependences as
the dynamics of tube inflation is different in both cases.
The osmotic shock limit is most likely not physiologi-
cally accessible to ion pumps. Crucially, one can see in
Fig. 3 that the instability length scale is set by dynamical
parameters, most importantly the ratio of the viscosity
and pumping time-scales. Varying τη/τpump has the ef-
fect of changing the time-scale over which the modes go
unstable. It is fortuitous that the dominant wavelength
does not depend strongly on the pumping rate (see S.I. -
Fig.S7), the parameter we can estimate least accurately.
This suggests a robustness to the wavelength selection
that may have important implications for the CVC's bi-
ological function. In the physiologically accessible range
of parameters for pumping and permeation, this length
scale is much larger than the asymptotic limit for either
the Rayleigh instability or the osmotic shock instability.
We have developed a model for a water-permeable
membrane containing uni-directional ion pumps. Hydro-
osmotic instabilities realised in cells may be expected to
usually lie in this class. Deriving dynamical equations
for a membrane tube we identify an instability driven by
this osmotic imbalance. This has a natural wavelength
that is set by dynamical parameters and is significantly
longer than a Rayleigh or Pearling instability but is of the
same order as seen in the CVC radial arm. We speculate
that this instability may provide a mechanism for bio-
genesis of the CV from a featureless active tube: bulges
in the radial arm are similar in size to the main CV. We
will further address the question of this oganellogenesis
FIG. 2. Plot of the distibution of mode amplitude (cid:112)(cid:104)¯u2
reaches (cid:112)(cid:104)¯u2
q(cid:105)
against scaled wavenumber q = qReq for t = 2.0 (solid), 2.04
(dashed) and 2.08 (dash-dotted, the time when the first mode
q(cid:105) = Req), τη/τpump ∼ 10−6. Req = 25nm, γ =
10−4N m−1 and ρI (t = 0) = 3.0 × 108µm−3
Eq. 16, but is more easily done by considering the time-
dependent growth rate Eq. 12 (graphically presented in
the S.I. - Fig.S2). Indeed, at the time t = tfinal, the dom-
inant mode q whose amplitude reaches (cid:112)(cid:104)¯uq2(cid:105) = Req
√
is very close in value to the fastest growing mode (the
peak of the instantaneous growth rate) at that particu-
lar time, written q∗, which can be derived analytically as
a function of the tube radius from Eq. (12). As a result
of the quasi-static driving of the instability by the ion
pumps, the final radius is always only marginally above
the critical radius
3Req, see S.I. This is the main fac-
tor contributing to the long wavelength/small q instabil-
ity. In this regime, the fastest growing mode is given by
q∗ ≡ Reqq∗ =
to leading or-
der, see S.I. While a qualitatively similar regime exists
for tension driven instabilities, it is only valid very close
to the instability threshold and its observation would
require a very precise tuning of the tension. Far from
threshold, the Rayleigh or Pearling instability shows a
universal relationship q∗ ∼ 0.6Req [21 -- 23, 26].
(cid:16) R(tfinal) − √
(cid:17)1/2
1√
2(3)1/4
3
A related instability is that of a membrane tube un-
der osmotic shock (see S.I.), for which one finds the most
unstable mode to be q∗ ∼ 0.2. The difference between
the Rayleigh and osmotic shock instabilities is due to the
growth rate having a different response when driven by
a volume change compared to surface tension, see S.I.
for details. The constant volume (Rayleigh) instabil-
ity might be of limited relevance for the morphological
changes of cellular membrane tubes, as cellular mem-
branes typically contains a host of membrane channels,
including water channels, which allow fairly rapid water
transport across the membrane. The osmotic instabil-
00.050.10.150.20.250.3~q10-310-210-1100101qhj7uqj2i=Req~t=2~t=2:04~t=2:08=2==7=2==pump0.0510-20.110-20.15^q10-30.210-40.2510-410-610-510-85
[23] R. Bar-Ziv, T. Tlusty, and E. Moses, Physical Review
Letters 79, 1158 (1997).
[24] K. L. Gurin, V. V. Lebedev, and A. R. Muratov, Journal
of Experimental and Theoretical Physics 83, 321 (1996).
[25] P. Nelson, T. Powers, and U. Seifert, Physical Review
Letters 74, 3384 (1995).
[26] G. Boedec, M. Jaeger, and M. Leonetti, Journal of Fluid
Mechanics 743, 262279 (2014).
[27] P. A. Pullarkat, P. Dommersnes, P. Fern´andez, J. F.
and A. Ott, Physical Review Letters 96, 1
Joanny,
(2006).
[28] W. Helfrich, Zeitschrift fr Naturforschung C 28, 693
(1973).
[29] R. Phillips, J. Kondev, J. Theriot, and H. Garcia, Physi-
cal Biology of The Cell, 2nd ed. (Garland Science, 2012).
[30] S. A. Safran, Statistical Thermodynamics of Surfaces, In-
terfaces, and Membranes, 1st ed. (Westview Press, 2003).
[31] D. Nelson, T. Piran, S. Weinberg, M. E. Fisher,
S. Leibler, D. Andelman, Y. Kantor, F. David,
D. Bertrand, L. Radzihovsky, M. J. Bowick, D. Kroll,
and G. Gompper, Statistical Mechanics of Membranes
and Surfaces, 2nd ed. (World Scientific, 2008).
[32] M. Chabanon, J. C. Ho, B. Liedberg, A. N. Parikh, and
P. Rangamani, Biophysical Journal 112, 1682 (2017).
[33] J. Zimmerberg and M. M. Kozlov, Nature reviews. Molec-
ular cell biology 7, 9 (2006).
[34] G. Koster, M. VanDuijn, B. Hofs, and M. Dogterom,
Proceedings of the National Academy of Sciences of the
United States of America 100, 15583 (2003).
[35] M. B. Jackson, Molecular and Cellular Biolophysics, 1st
ed. (Cambridge University Press, 2006).
[36] R. D. Allen and A. K. Fok, J. Protozool 35, 63 (1988).
[37] T. Tani, R. Allen, and Y. Naitoh, J. Exp. Biol. 203, 239
(2000).
[38] K. Olbrich, W. Rawicz, D. Needham,
and E. Evans,
Biophysical Journal 79, 321 (2016).
[39] J. D. Murray, Asymptotic Analysis, 1st ed. (Springer,
1984).
[40] Lord Rayleigh, Philos. Mag. 34, 145 (1892).
in future work.
Thanks to J. Prost and P. Bassereau (Paris), M. Polin
(Warwick) and R. G. Morris (Bangalore) for interest-
ing discussions and insight. Thanks to the reviewers
for helpful comments and for alerting us to references
[20, 26]. S. C. Al-Izzi would like to acknowledge fund-
ing supporting this work from EPSRC under grant num-
ber EP/L015374/1, CDT in Mathematics for Real-World
Systems.
[1] C. Mullins, The Biogenesis of Cellular Organelles, 1st ed.
(Springer, 2005).
[2] Y.-H. M. Chan and W. F. Marshall, Science 337, 1186
(2012).
[3] B. Alberts, A. Jihnson, J. Lewis, M. Raff, K. Roberts,
and P. Walter, Molecular Biology of The Cell, 5th ed.
(Garland Science, 2008).
[4] R. Heald and O. Cohen-Fix, Current Opinion in Cell Bi-
ology 26, 79 (2014), cell architecture.
[5] Y. Shibata, J. Hu, M. M. Kozlov, and T. A. Rapoport,
Annual Review of Cell and Developmental Biology 25,
329 (2009).
[6] U. Jeleri and N. S. Gov, Physical Biology 12, 066022
(2015).
[7] D. Gonzalez-Rodriguez,
S. Sart, A. Babataheri,
D. Tareste, A. I. Barakat, C. Clanet, and J. Husson,
Phys. Rev. Lett. 115, 088102 (2015).
[8] A. Yamada, A. Mamane, J. Lee-Tin-Wah, A. D. Cicco,
D. Levy, J.-F. Joanny, E. Coudrier, and P. Bassereau,
Nature communications 5, 3624 (2014).
[9] D. J. Patterson, Biological Reviews 55, 1 (1980).
[10] R. D. Allen, BioEssays 22, 1035 (2000).
[11] K. Komsic-Buchmann, L. Wostehoff, and B. Becker, Eu-
karyotic Cell 13, 1421 (2014).
[12] C. Stock, H. K. Grønlien, R. D. Allen, and Y. Naitoh,
Journal of cell science 115, 2339 (2002).
[13] Y. Naitoh, T. Tominaga, M. Ishida, A. K. Fok, M. Ai-
hara, and R. D. Allen, The Journal of experimental bi-
ology 200, 713 (1997).
[14] R. Docampo, V. Jimenez, N. Lander, Z.-H. Li,
and
S. Niyogi, International Review of Cell and Molecu-
lar Biology, Vol. 305 (Elsevier, 2013) pp. 69 -- 113, dOI:
10.1016/B978-0-12-407695-2.00002-0.
[15] J. Heuser, Q. Zhu, and M. Clarke, Journal of Cell Biology
121, 1311 (1993).
[16] T. Nishi and M. Forgac, Nature Reviews Molecular Cell
Biology 3, 94 (2002).
[17] A. K. Fok, M. S. Aihara, M. Ishida, K. V. Nolta, T. L.
Steck, and R. D. Allen, Journal of cell science 108 ( Pt
1), 3163 (1995).
[18] T. Tominaga, R. Allen, and Y. Naitoh, The Journal of
experimental biology 201, 451 (1998).
[19] Rayleigh, The London, Edinburgh, and Dublin Philo-
sophical Magazine and Journal of Science 34, 145 (1892).
[20] S. Tomotika, Proceedings of the Royal Society of London.
Series A, Mathematical and Physical Sciences 150, 322
(1935).
[21] T. R. Powers and R. E. Goldstein, Physical Review Let-
ters 78, 2555 (1997).
[22] R. Bar-Ziv and E. Moses, Physical Review Letters 73,
1392 (1994).
Supplementary Information
Differential geometry of the membrane
6
(cid:90)
F =
dA
S
(cid:16) κ
2
(cid:17)
This manifold has mean curvature H and constant surface tension γ. The free energy for such a membrane is
(2H)2 + γ
,
(S1)
where dA is the area element on S and κ is the bending rigidity.
For the membrane tubes in which we are interested we parametrise the bilayer as an embedding in R3. Utilising
the cylindrical symmetry of the membrane tube we write this as a surface of revolution about the z axis with radius
r(z, t). This means that we will only consider squeezing (peristaltic) modes in our analysis. In Cartesian coordinates
this surface is parametrised by the vector (cid:126)R = (r cos θ, r sin θ, z), i.e. by the normal cylindrical polar coordinates.
From this we can induce covariant coordinates on the manifold as
(cid:126)e1 =
(cid:126)e2 =
∂ (cid:126)R
∂θ
∂ (cid:126)R
∂z
= (−r sin θ, r cos θ, 0)
= (∂zr cos θ, ∂zr sin θ, 1) .
This allows for the definition of a Riemannian metric as
Hence the metric and its inverse are
g =
gij = (cid:126)ei · (cid:126)ej
0
(cid:21)
(cid:20)r2
0 1 + (∂zr)2
i, j = {1, 2},
for
(cid:20) 1
r2
0
g−1 =
,
0
1
1+(∂zr)2
(S2)
(S3)
(S4)
(S5)
(cid:21)
.
To find the curvature of S we need to know how the normal vector, (cid:126)n, to the surface S varies. We can write this
normal vector as
(cid:126)n =
(cid:126)e1 × (cid:126)e2
(cid:126)e1 × (cid:126)e2 =
1(cid:113)
1 + (∂zr)2
(cos θ, sin θ,−∂zr) .
(S6)
From this we can find the second fundamental form bij = (cid:126)n · (cid:126)ei,j where the comma denotes a partial derivative.
Taking the determinant and trace of
b j
i =
1(cid:113)
we find the mean and Gaussian curvatures
2H =
K =
r
1 + (∂zr)2
(cid:32)
1(cid:113)
1 + (∂zr)2(cid:17)2 .
(cid:16)
1 + (∂zr)2
−∂zzr
(cid:34)−1
r
0
(cid:35)
,
0
∂zzr
1+(∂zr)2
(cid:33)
∂zzr
1 + (∂zr)2 − 1
r
(S7)
(S8)
(S9)
If we consider the case where we are only interested in membranes that do not change topology then we can apply
Gauss-Bonnet theorem to the free energy term, integrating out the constant contribution of the Gaussian curvature.
Dynamics of radial growth due to water permeation
Approximate solution for slow pumping
Fig.S1 shows the agreement between the asymptotic solution to the radial dynamics: R(t) = (1 + t/τpump)1/2
(Eq.10 - main text) and the full numerical solution of Eq.9 (main text).
7
FIG. S1. Left: plot showing approximate solution (dots) and full numerical solution (solid line). Right: plot showing the
absolute error between the approximate solution and numerical solution.
We can find the radius R(q) at which the mode q first goes unstable by finding the zero of the α(q) polynomial,
Eq.4 (main text), defining R(q) =
√
3Req + δR(q). For the small q limit and assuming δR(q)
Req
is small we find
√
δR(q)
Req
≈
3 (Reqq)2 .
(S10)
Using Eq.(S10) with the approximate solution for R(t) gives a formula for the time the mode q first goes unstable
(S11)
where q = qReq.
q ≈ τpump
t∗
(cid:0)2 + 6q2(cid:1)
Case of an osmotic shock
We can consider a tube with a fast-acting tension reservoir (something similar to the smooth spongiome), undergoing
osmotic shock. It is interesting to understand the dominant wavelength selection in such a case as the system may be
easier to implement in vitro than systems involving unidirectional ion pumps. If the radial expansion of the membrane
is driven by a hypo-osmotic shock, the radial dynamics are governed by the following growth equation
(cid:18) 1
R2
d R
dt
=
1
R
(cid:18)
(cid:19)(cid:18) 1
(cid:19)(cid:19)
+
− 1
1 +
γ
R
∆ρ
ρ0
and ∆ρ = ρ0 − ρshock is the change in ionic density of the outside
(S12)
R2
where t = t
τµ
medium due to osmotic shock. Note that the normalisation chosen here is different from the one used in the text.
, R = R
Req
, τµ = Req
kB T Reqρ0
, γ =
µkB T ρ0
γ
0123451.01.21.41.61.8t~R~012345-0.0006-0.0005-0.0004-0.0003-0.0002-0.00010.0000t~ΔR~RSolution to the stokes equations
8
The equations governing the bulk flow are just the standard inertia free fluids equations for velocity field (cid:126)v. These
are the continuity equation for incompressible flow
and the Stokes equation
(cid:126)∇ · (cid:126)v = 0
(cid:126)∇P = η∇2(cid:126)v
where P is the hydrodynamic pressure and η the viscosity. The system has the following boundary conditions
vrr=R = u + vp
vzr=R = 0
(S13)
(S14)
(S15)
(S16)
where vp is the permeation velocity, the second boundary condition is justified by any local area change in the
membrane coming from exchange with the tension reservoir, the surrounding sponge phase. Hence there is no lateral
membrane flow (at least at this order).
If we write the velocity field in terms of a stream function ψ as
the continuity equation is automatically satisfied, and the Stokes equations can be solved to give
(cid:126)v =
1
r
(∂zψ(cid:126)er − ∂rψ(cid:126)ez)
(cid:40)(cid:80)
(cid:80)
ψ =
q A1qrI1(qr) + B1(qr)2I0(qr)
q A2qrK1(qr) + B2(qr)2K0(qr)
r < R
r > R
Eq.(S21) can be used to describe the dynamical instability of a membrane tube subjected to different driving
mechanisms; an increase of membrane tension (Rayleigh instability), an osmotic shock, or the slow active pumping
mechanism we are primarily interested in. In the limit qR (cid:28) 1 this gives
2µR(t)
(cid:18) q2R(t)
(cid:19)
¯uq = −αL(q)
+
8η
R3
eq
¯uq
in the interior of the tube, where A1,2 and B1,2 are found from the boundary conditions. Iν(x) and Kν(x) are modified
Bessel functions of the first and second kind respectively.
From here we use the equation vp = µ(∆P )r=R, where ∆Pr=R is the hydrodynamic pressure jump across the
tube membrane, and use the solution of the interior and exterior hydrodynamic pressure from the Stokes equations
to find a value of vp. In Fourier space this gives
¯vp =
¯uq
2qηχ(q)µ − 1
1
where
χ(q) =
I0 (I0 − 1)
0 − 2I0I1 − qRI 2
1
qRI 2
−
K 2
0
qRK 2
0 + 2K0K1 − qRK 2
1
The force balance equation at the membrane reads
(P − 2η∂rvr)r=R = f
.
(S19)
(S20)
where f is the force required to displace the membrane to u and can be found from the free energy. Substituting the
velocity and pressure fields into this gives the dynamic equation for the modes ¯uq
with the shorthand Iν = Iν(qR) and Kν = Kν(qR)[24]. The elastic response function αL(q) is obtained by replacing
the pressure P by the Laplace presure PL = γ/R − κ/(2R3) in Eq.4 of the main text:
¯uq = − αL (q)
2ηR
1
(1 − 2qµχ(q)) ¯uq
X(q) =
qR (I 2
0 ) + 2I1I0
+
qR (K 2
X (q)
I0 (qRI0 − I1)
1 − I 2
(cid:18)
αL(q) =
κ
R2
(qR)4 − 1
2
(qR)2 +
3
2
K0 (qRK0 − K1)
0 ) + 2K1K0
1 − K 2
(cid:19)
+ γ(cid:0)(qR)2 − 1(cid:1)
(S17)
(S18)
(S21)
(S22)
(S23)
(S24)
9
FIG. S2. Location of the peak of the growth rate (q∗ ≡ Reqq∗) for a tube under constant tension, as a function of the tube
radius. The initial tube radius Req corresponds to the equilibrium radius of a tube under zero pressure.
Mode growth rates
¯uq
¯uq
We define the instantaneous growth rate G(q) =
from Eq.(S21). This growth rate shows a peak as a function of
q. The location of the peak depends on how the instability is driven. Starting with a stable tube under zero pressure
with radius R0 and membrane tension γ0, the instability can be driven by an increase of tension γ > γ∗ = 3γ0 at
constant volume (Rayleigh instability), or by an increase in volume (or radius) R > R∗ =
3R0 at constant tension
(Osmotic instability). In the former case, and in the limit γ (cid:29) γ∗, the growth rate reaches a universal shape with
a peak at R0q∗ (cid:39) 0.6. The most unstable wavelength is thus entirely set by the initial tube geometry (its radius
R0). In the latter, the peak of the growth rate depends on the time-dependent radius and does not reach any sort of
universal behaviour. In fact the location of the peak is a non-monotonic function of the radius, first increasing, then
decreasing with increasing radius. Its largest possible value is R0q∗ (cid:39) 0.2 and occurs for R (cid:39) 2.35R0, see Fig.S2.
√
As the fastest growing mode changes in time, it is the cumulative growth that is important. This means we must
integrate the growth of each q mode over time, accounting for fluctuations, as discussed in the main text. However,
q(cid:105) = Req at t = tfinal) is close
due to the exponential growth the dominant q-mode (the one that first satisfies
to the fastest growing q-mode at that particular time. The latter can be expressed in terms of δ R(tfinal) = δR
=
R(tfinal) − √
Req
3, Fig.S3. It is important to note that whilst the growth rate relation does give a good approximation to
(cid:113)(cid:104)¯u2
the dominant wavelength, there is a difference due to the history encoded in the full dynamical description.
The peak of the growth rate relation can be found analytically ( dG
dq q∗ = 0), and in the small q limit is
(cid:114)
−1 − 32 ηµ
Req
+ 1
R2 +
q∗ =
)+ R4(1+32 ηµ
Req
(−1+32 ηµ
Req
))
R2
(S25)
(cid:113)−17+4 R2(1+8 ηµ
Req
√
6
to leading order, in the µ → 0 limit, this can be expressed as q∗ =
main text.
(δ R(tfinal))1/2
√
2(3)1/4
, which is the expression given in the
The growth rate relation is quantitatively different from a Rayleigh instability due to the driving mechanism.
The functional dependence of the growth rate relation depends on the polynomial αL(q) describing the membrane
mechanics in q space (Eq.(S23)). The Rayleigh instability is driven by a surface tension γ > 3κ
at constant volume
2R2
0
(R(t) = R0), so that the magnitude of the q4 term in Eq.(S23) doesn't change.
In the case of osmotic pressure
2γ . This increases
however, the instability is driven by a change in volume caused by the osmotic pressure, i.e. R >
the prefactor to the q4 term which means that the higher q modes are stabilised compared to the Rayleigh case. This
means that the dominant wavelength is skewed towards smaller q, Fig.S4.
(cid:113) 3κ
Inserting the time-dependent solution of Eq.(S12) in the growth equation Eq.(S21) (including thermal noise, as in
Eq.13 of the main text) gives access to the evolution of the amplitude of the different modes. The exact value of
Osmotic shock
2.02.53.03.54.00.000.050.100.150.20Rq*10
FIG. S3. Dominant wave-number squared, q2, plotted against final radius minus critical radius δ R(tfinal) = R(tfinal) − √
3, the
solid line corresponds to the peak of the growth rate as a function of wavenumber and points represent the peak found by
numerically solving the full dynamics.
FIG. S4. Normalized growth rate relation for a membrane tube undergoing a Rayleigh instability (R0 ∼ 10−2µm, µ = 0,
κ = 10kBT , γ = 89γ0, where γ0 = κ/(2R2
0)) or responding to an osmotic shock under constant membrane tension (R(t = 0) =
Req = 10−2µm, µ = 10−4µm Pa−1s−1, κ = 10kBT , R(t) = 2.35). These parameters are chosen such that they illustrate the
growth rate relations in the high tension limit for the Rayleigh instability (blue curve), or correspond to the maximal peak
wavelength in the case of osmotic shock (orange curve). The dispersion relation for the Rayleigh instability is obtained from
Eq.(S21), with constant radius and the limit µ → 0. For comparison the typical growth rate for physiological parameters in
the case of slow pumping (with R =
3 + 0.05) is also shown (green curve).
√
FIG. S5. Surface plot showing the dominant wave-number of an instability driven by osmotic shock when varying permeation
time-scale, τµ, and shock magnitude ∆ρ/ρ0.
the dominant q depends on the permeability µ (or the time-scale τµ) and the magnitude of the shock ∆ρ/ρ0. A 3D
plot of how this varies is shown in Fig.S5. Comparison with the behaviour that arises in the presence of ion pumps
0.10.20.30.4δR~(tfinal)0.010.020.030.04q~*20.00.20.40.60.81.00.00.20.40.60.81.0qG(q)=2==7";=;00.110-20.120.141020.16^q0.1810-40.20.2210110-610011
FIG. S6. Plot of dominant wavenumber, q, against cutoff criterion, C (see text). Crucially any dependence on C is very weak.
The values of τpump and τµ have been chosen to correspond with the four corners of the surface plot Fig. 3 in the main text.
All times are in units of seconds.
(Fig. 3 - main text) shows that the peak value of the dominant mode is the same in both case, and corresponds by
the peak of Fig.S2. This peak occurs for fast pumping (τη/τµ > 10−2 - Fig. 3 - main text) or for strong osmotic shock
(∆ρ/ρ0 > 10 - Fig.S5), showing that these two situations are somewhat similar. However the details are different due
to the different dynamics of tube inflation in both cases.
The drop off in dominant wavelength of the osmotic shock instability when permeability and shock magnitude are
very large is caused by the decrease of the peak of the growth rate relation at very large radii (Fig.S2). This happens
because of a decrease in the contribution of the bending rigidity to the energy at large radii and small q. The surface
tension contribution to the energy remains, hence the instability starts to be dominated by surface tension. The only
contribution of the bending terms is to increasingly stabilise the larger values of q, thus pushing the peak wavelength
to lower q. Interestingly the bending rigidity in this limit acts in a qualitatively similar manner to a large difference
in viscosities discussed in the original fluid jet papers [19, 20].
Defining the dominant wavelength
Defining the dominante wavelength of a time dependent growth rate is in general a difficult task; as the peak of the
dispersion relation is time dependent we must instead consider the full growth history of each mode. We define the
dominant mode at linear order to be the first one to have (cid:104)¯uq2(cid:105) = CR2
eq where C = 1. It is therefore a sensible thing
to check that the chosen value of the cutoff, C, has a minimal effect on our results, i.e. the dominant wavelength at
linear order should be constant for C ∼ 1. Plotting q∗ against C, Fig.S6, shows a weak logarithmic dependence of the
dominant wavenumber on C. The only pronouced effect for a cutoff around linear (C ∼ 1) order might be to shift
the values in the fast pumping limit by < 5%, the values for physiological parameters remain virtually unaffected.
Weak dependence of dominant wavelength on the pumping rate in the physiological range
The asymptotic solution presented in the main paper is valid for parameter estimates consistent with the CVC. It
is of interest to see how the wavelength of the instability varies with pumping rate in this limit. The wavelength of the
0.10.51510501000.000.050.100.150.20Cq*τpump=10,τμ=10-4τpump=10-4,τμ=10-4τpump=10,τμ=5×10-2τpump=10-4,τμ=5×10-2Printed by Wolfram Mathematica Student Editioninstability varies with pumping rate but very weakly (slower than logarithmically). The wavelength for time-scales
consistent with the CV pumping is λ ∼ 1µm which is of the correct order of magnitude for the CV and much larger
than the tube radius. The weak dependence of the wavelength on the pumping provides a robust mechanism of size
regulation, Fig.S7.
12
FIG. S7. Plot of dominant length, λ, of instability against ratio of viscous to pumping time-scales τη/τpump for the asymptotic
solution found in the main paper (Eq. 10 - main text). Here τη/τµ = 10−4. This plot is essential a cross-section of Fig. 3 in
the main paper for τη/τµ = 10−4, but plotting wavelength instead of wavenumber q∗.
Note on numerical implementation
All the numerics shown in Fig. 2 and Fig. 3 of the main paper are implemented using a discrete Fourier transform, as
such the autocorrolation function, (cid:104)¯uq2(cid:105) has units of [Length]2, this choice of implementation is used to simplify the
criterion for the fully developed instability. The longest mode in real space is chosen to be 104Req, this corresponds
to a small enough spacing for the q space to approximate a continuum.
10-710-610-510-410-3=2==pump304050607080901006=Req |
1907.11691 | 1 | 1907 | 2019-07-26T17:47:12 | Flocking in complex environments -- attention trade-offs in collective information processing | [
"physics.bio-ph",
"physics.soc-ph"
] | The ability of biological and artificial collectives to outperform solitary individuals in a wide variety of tasks depends crucially on the efficient processing of social and environmental information at the level of the collective. Here, we model collective behavior in complex environments with many potentially distracting cues. Counter-intuitively, large-scale coordination in such environments can be maximized by strongly limiting the cognitive capacity of individuals, where due to self-organized dynamics the collective self-isolates from disrupting information. We observe a fundamental trade-off between coordination and collective responsiveness to environmental cues. Our results offer important insights into possible evolutionary trade-offs in collective behavior in biology and suggests novel principles for design of artificial swarms exploiting attentional bottlenecks. | physics.bio-ph | physics | Flocking in complex environments -- attention trade-offs in
collective information processing
Parisa Rahmani,1, 2 Fernando Peruani,3 and Pawel Romanczuk2, 4, ∗
1Department of Physics, Institute for Advanced Studies
in Basic Sciences (IASBS), Zanjan 45137-66731, Iran
2Institute for Theoretical Biology, Department of Biology,
Humboldt Universitat zu Berlin, Germany
3Universit´e Cote d'Azur, Laboratoire J.A. Dieudonn´e,
UMR 7351 CNRS, Parc Valrose, F-06108 Nice Cedex 02, France
4Bernstein Center for Computational Neuroscience, Berlin, Germany
(Dated: July 29, 2019)
Abstract
The ability of biological and artificial collectives to outperform solitary individuals in a wide va-
riety of tasks depends crucially on the efficient processing of social and environmental information
at the level of the collective. Here, we model collective behavior in complex environments with
many potentially distracting cues. Counter-intuitively, large-scale coordination in such environ-
ments can be maximized by strongly limiting the cognitive capacity of individuals, where due to
self-organized dynamics the collective self-isolates from disrupting information. We observe a fun-
damental trade-off between coordination and collective responsiveness to environmental cues. Our
results offer important insights into possible evolutionary trade-offs in collective behavior in biology
and suggests novel principles for design of artificial swarms exploiting attentional bottlenecks.
9
1
0
2
l
u
J
6
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
1
9
6
1
1
.
7
0
9
1
:
v
i
X
r
a
∗ [email protected]
1
INTRODUCTION
Consensus formation, coordination and collective response to environmental cues are im-
portant aspects in collective behavior of many interacting agents in biology, physics, robotics
and computational social science. The understanding of these processes is fundamental for
a better comprehension of collective intelligence that confers groups the ability to solve
problems collectively using strategies which are beyond the reach of single individuals [1].
Crucial for understanding benefits of collective behavior, is the understanding of the mech-
anisms underlying collective information processing: How new information is acquired, how
information is shared and combined within the collective, and how the collective deals with
conflicting information are among the most compelling and elusive questions on the self-
organization of collectives. Generic flocking models (see e.g.
[2 -- 5]) allow to study the
interplay between emergent collective behaviors and collective information processing in a
dynamical system setting. For example, using flocking models it has been demonstrated, that
only a small fraction of informed individuals is sufficient to accurately guide large collectives
[6 -- 8]. It was also shown that groups are able to collectively track dynamic environmental
gradients not detectable by individuals [4, 9], or that groups can make efficient consensus
decisions in conflict situations without any implicit knowledge about the majority-minority
relationships [6, 10]. Only recently, predictions of such models on fundamental decision bi-
furcations in spatial movement decisions have been confirmed in the collective migration of
baboon groups [11].
A fundamental aspect of the self-organization of collectives -- from collective decision-
making to consensus formation, including coordinated movements -- is that individuals are
limited in terms of perception and cognition. Without direct access to the state of the whole
group they must rely on local information [12 -- 15]. The emergent collective patterns in agent-
based models have been shown to depend strongly on the field of view of individuals [16,
17], and more generally on what local information individuals pay attention to [18, 19].
Furthermore, even for a strongly limited field of view, the sensory input of individual agents
may contain a large number of social and non-social cues. However, social interaction in
animal groups appear to be restricted to a rather low number of neighbors [20, 21]. On
the one hand, this suggests additional cognitive constraints on the processing of available
sensory information, which is also in-line with a wide range of experimental results on limited
capacity for visual tracking of multiple objects in animals and humans [22 -- 26]. On the other
hand, it has been shown in generic flocking models that the emergent, large scale collective
behavior depends strongly on how many neighbors a given individual can pay attention
to [27 -- 29]. However, to our knowledge the explicit role of cognitive constraints on collective
information processing has not been systematically explored.
Up-to-date most theoretical and empirical research focused on information sharing and
collective decision making, in idealized, laboratory-like environments, including the works
mentioned above, discussing cognitive and sensorial limitations (see e.g.
[6, 10, 30, 31]).
However, collectives in realistic scenarios need to cope with complex environments with a
large number of potentially informative or distracting environmental cues (see Fig. 1a and
[32]). Whereas recently it was shown using minimal flocking models that self-organized,
collective behaviors are strongly affected by complex environments [33, 34], it remains open
how complex environments impact collective information processing, which is the main ques-
tion we focus on here. More specifically, we investigate the emergence of collective behaviors
in a generic model of socially interacting agents in a complex environment containing many
2
FIG. 1. a: Schematic visualisation of attention trade-off in collective behavior in complex envi-
ronments. The focal individual can only pay attention to k = 3 nearest objects -- other agents or
non-social environmental features -- simultaneously. b: Visualization of different situations that
may occur in the model. The arrows indicate the velocity vectors (cid:126)v of the different agents. The
small black circles indicate the location of danger sites l and l(cid:48) with their repulsion zones shown
in blue. Agent i (red) reacts to the danger site (DS) l as two conditions are met simultaneously:
DS l is in i's kNO, and agent i is also within the corresponding repulsion zone. Agent j (magenta)
does not react to DS l(cid:48) since it's attention slot is already filled with three other agents (one blue
and two gray). Agent k (blue) perceives DS l(cid:48) but does not react to it, because it is outside of the
repulsion zone. It only reacts to two other neighbors (gray and magenta) and aligns with them.
potentially dangerous sites.
Individuals try to avoid these sites, while at the same time
trying to coordinate with their neighbors. In addition, some informed individuals have also
private information on a global preferred direction of migration, which may be in conflict
with the local environmental cues. Our main aim is to explore how the cognitive, and/or
sensory constraints of the individuals affect group-level coordination, information exchange
and collective response to environmental cues.
Our results show that in heterogeneous environments, strongly limited attention capabil-
ity of individual agents results in higher accuracy with respect to large-scale coordination,
which is in stark contrast to previous results obtained in simple environments [28, 29]. This
is caused by a dynamical, spatial "echo chamber"-like effect, where individual attention
becomes saturated by social information and non-social cues are largely ignored. However,
if these non-social cues provide important information about potential environmental dan-
gers, the emergent dynamical "echo chambers" become strongly detrimental to the ability of
the collective to safely navigate the environment. Note that information exchange through
social interactions is typically believed to be beneficial for the collective [9, 35]. Here, our
analysis shows that below a critical threshold in attention capacity, groups perform worse
at acquiring new information about the environment than non-interacting agents. This is
due to the emergent self-isolation from environmental cues, which is exactly what facilitates
group coordination in complex environments. Our findings not only suggest a fundamental
trade-off in collective behavior in natural systems, but also provide important insights for
the design of communication in artificial, distributed systems, such as robotic swarms.
3
l΄lφiαi,l⃗vi⃗vj⃗vkabRESULTS
Model. We consider a flocking model consisting of N agents moving in a two-dimensional
environment of size L × L with periodic boundary conditions. In addition, we assume the
presence of Ninf informed individuals with private information about a preferred direction
of motion up. The informed fraction of the collective is Rinf = Ninf/N .
The environment contains non-social cues, which represent features of the environment
that solitary agents in general try to avoid, as they, for example, signal potential dangers.
We will refer to these disrupting cues as "danger sites" (or distraction sites) in short as
DS. Each DS is surrounded by an effective repulsion zone of radius r = 1. The environment
contains NDS randomly distributed DSs at fixed positions. In particular at high DS densities
ρDS = NDS/L2, the corresponding repulsion zones may overlap (see Fig. 2a, b). We note
that the agents and the danger sites (DS) are assumed to be point-like, as in many agent-
based models for collective movement, e.g. Vicsek-type models [3]. This corresponds to the
scenario where the sensory ranges are large compared to the physical size of moving agents
and DSs.
Based on experimental observations, it was suggested that animals interact with a limited
number of conspecifics [14, 20, 21]. Motivated by these findings, and with intention of
explicitly studying the impact of limited attention in a generic flocking model, we assume
that each agent can pay attention only to k nearest objects (kN O) in its vicinity, irrespective
whether it is another agent (social cue) or a DS (non-social cue). Thus, k can be interpreted
as the number of available attention slots for each agent. The parameter k quantifies the
individual attention capacity.
Each agent moves with a constant speed v0 and reacts to neighbors and DSs through
corresponding changes in its direction of motion ui = (cos ϕi, sin ϕi)T defined by a polar
orientation angle ϕi. All agents - informed and uninformed - turn away from DSs when
two conditions are met simultaneously: 1) the agent perceives the DS -- that means that
the DS in question, say l, is within its k nearest objects -- and 2) the distance between
the agent and DSs (dil = (cid:126)xi − (cid:126)xl) is smaller than the radius of its repulsion zone. In all
other cases, individuals ignore the DSs and coordinate their motion with other individuals
within their kNO by aligning their direction of motion with the average direction of their
neighbors. Informed individuals exhibit an additional bias to orient towards the preferred
direction of motion that we denote up. Throughout this work, the preferred direction of
motion of informed individuals will be along the x-axis: up = (1, 0)T . We emphasize that
the response to DSs, once detected, dominates all other behavioral responses of individuals,
whether informed or uninformed. The specific formulation of the mathematical model in
terms of stochastic differential equations, together with the parameters used, are given in
Methods.
The finite attention capacity to k nearest objects leads to a natural competition between
social and non-social cues: If the k nearest objects of the focal agent are other agents, it
will be not capable to "sense" a DS l even if dil < 1 (see Fig. 1b). Note that in the
case of vanishing density of DSs ρDS → 0 (homogeneous environment), the model reduces
to a simple flocking model with so-called metric-free alignment interaction with k-nearest
neighbors (see e.g. [20, 28, 36]) with the additional feature of informed individuals. If instead
of topological, metric interaction are used, then the model reduces in the limit of ρDS → 0
and Rinf = 0 to a Vicsek-type alignment model [3], which has been extensively discussed
in [37, 38]. For Rinf > 0 it is closely related to the model explored in [6], while for Rinf = 0
4
and ρDS > 0 it reduces to the model studied in [33, 34].
Interaction Networks & Collective Accuracy. In order to quantify the emergent
collective dynamics and study the effect of varying attention capability, we have performed
systematically numerical simulations of the above model for varying attention limit k, DS
density ρDS, and the ratio of informed individuals Rinf (see Methods for details). By ne-
glecting the directed nature of inter-individual links, the entire agent system can be viewed
as a time-dependent, undirected interaction network. For all DS densities, we obtain the
same qualitative picture in the stationary state: For low k, we observe strongly fragmented
dynamical networks, which at given time t are characterized by a large number of small,
disconnected sub-groups (see Fig. 2a and Supp. Fig. S1). Each such cluster corresponds to
an isolated connected component. These components are not static: We observe continuous
fission-fusion of clusters over time due to randomness in individual motion and interactions
with DSs (see Movie S1). By increasing the attention limit k, we observe a fast decrease
in the number of disconnected clusters that results in an increase in average cluster size
(see Fig. 2b and Supp. Fig. S2a). Eventually, by increasing k above a critical value, we
can obtain fully connected networks with a single connected component. Correspondingly,
the average life-time of a connection between specific agents grows strongly with increasing
attention limit k, whereas an increase in DS density ρDS reduces the life-time of individual
edges in the network (see Supp. Fig. S2b). In general, as one would intuitively expect,
the network of social interactions becomes more tightly connected with increasing attention
capacity k.
We measure the accuracy of the collective migration through the average agreement
between the heading uj of individuals and the preferred direction of motion up:
(cid:42)
1
N
(cid:88)
j
C =
uj · up
(cid:43)
(1)
with (cid:104)·(cid:105) indicating the temporal average in the stationary state. If all agents move perfectly
along the preferred direction, which is available only to informed individuals, then we obtain
C = 1, whereas for disordered movement we observe C ≈ 0.
For collective behavior in homogeneous environments it has been shown that increasing
the connectivity of the interaction networks is beneficial for coordination [29, 30], which
is also in line with general results on synchronization in (dynamic) networks of oscillators
[39, 40]. In particular for few informed individuals, one intuitively expects that a strongly
connected information network ensures that information about the preferred direction of
motion diffuses more efficiently across large parts of the collective. Therefore, the natural
prediction would be that collective accuracy increases with increasing attention capacity
k. This is indeed the case in the limit of vanishing density of DSs ρDS = 0 (homogeneous
environment, see Fig. 3a), where we observe a monotonous increase in accuracy C with the
attention capacity k, for a fixed ratio of informed individuals: Whereas for k = 1, in order
to achieve an accuracy of C > 0.9, we require the majority of the collective to be informed
(Rinf ≥ 0.6), for k = 6 it is already sufficient to have a small fraction Rinf = 0.2 of informed
individuals to achieve the same level of collective accuracy.
The situation completely reverses for high DS densities (see Fig. 3b). For ρDS = 0.2 we
observe a monotonous decrease in the collective accuracy with increasing k for all values
Rinf > 0. In order to achieve a certain level of collective accuracy (e.g. C = 0.5) for larger k
we need a larger fraction of the system to be informed. In other words, stronger connected
5
FIG. 2. a,b: Examples of social interaction networks for k = 3 (a) and k = 12 (b) at ρDS =
0.25. The black symbols indicate socially interacting agents, whereas the red symbols indicate
agents responding to a DS. The lines indicate the (non-directed) interaction network. Filled circles
represent uninformed agents, empty circles indicate agents informed about the preferred direction
of migration. The DS positions shown by blue dots, are surrounded by a disc-like repulsion zones
(light blue). For clarity, only a portion of the respective simulation box is represented here, see
Supplementary Figure S1 for the full snapshots. c,d: In-out degree distributions for the emergent
social interaction networks for low attention limit k = 3 (c) and high attention limit k = 12 (d) at
low and high DS densities (ρDS = 0.05, and ρDS = 0.25). The vertical dashed lines are for visual
guidance to distinguish the subpopulations with Dout = 0 corresponding to agents responding to
DSs (left of the vertical line). At high density of DSs this distribution is clearly bimodal with
two peaks at Dout = 0 and Dout = k. By increasing DS density number of agents with Dout = 0
increases. These agents have a lower in-degree compared to non-responders, which contributes to
the self-isolation of the collective from environmental cues at low k values. Rinf = 0.1 in all the
panels in this figure.
flocks become more difficult to guide. Even more dramatically, for the largest attention
capacity investigated (k = 24) the maximal attainable average accuracy for a fully informed
system (Rinf = 1) is C ≈ 0.62. This is lower than the average collective accuracy of C ≈ 0.66
for collectives at minimal attention capacity (k = 1) with only a tiny fraction of informed
individuals Rinf ≈ 0.013 (1.3% of the entire collective). In fact, it appears that for k = 1,
the collective accuracy C versus Rinf depends only very weakly on the DS density, whereas
the accuracy for high k is massively decreased for all Rinf > 0.
Thus, in complex environments, strongly limited attention resulting in a very sparse,
fragmented - yet dynamic - interaction networks, turns out to be highly beneficial for global
6
acdbk= 3k= 12FIG. 3. a, b: Collective accuracy C of migration along the preferred direction versus the ratio of
informed individuals for different attention limits k, for environments with no danger sites ρDS = 0
(a), and environments with high DS density ρDS = 0.2 (b). The red arrows show the direction of
increasing k. c: Collective accuracy C versus attention limit k for different DS densities ρDS at
Rinf = 0.1.
consensus formation (see Fig. 3c). This counter-intuitive effect can be understood through
an analysis of information flows and how environmental information is processed by the
collective, for example through analysis of the (stationary) probability distribution of agents
having a particular combination of in and out-degree. The out-degree Dout quantifies how
many individuals a focal individual pays attention to, whereas the in-degree Din is the
number of others paying attention to the focal individual. Agents directly responding to a
DS have a social out-degree Dout = 0 - they ignore their neighbors. However, their neighbors
can still pay attention to them so their social in-degree Din can be larger than zero (see Supp.
Information and Fig. 2c,d), in this case they "broadcast" information on the DS through
their evasion behavior to others. Therefore, environmental cues affect collective behavior
directly through the individuals directly responding to a DS, as well as, indirectly through
information transmitted to other agents via social interactions. High consensus in complex
environments requires effective self-isolation of the collective from distracting environmental
cues. It can emerge due to two mechanisms: 1) The number of direct responders remains
small; 2) Their influence on others is weak due to a low Din, in particular compared to the
in-degree of non-responders. In our case both effects play a role: For low k, the fraction
of agents directly responding to DSs remains low even at very high ρDS. At low k, aligned
individuals move together in dense sub-groups, and even if one or more agents enter a
repulsion zone, there is a high probability that there are k neighbors closer than the DS,
which prevents the detection of the latter. In addition, the in-degree of agents responding
to DSs is on average significantly lower than that of agents not responding to the DSs (Fig.
2c,d). Agents evading a DS, have a high probability to move away from their neighbors,
which in turn decreases the probability that they will be within the kN O of others. Thus,
in particular for low k, the indirect response to DSs, mediated through social interactions,
is strongly inhibited. These emergent small, dense agent clusters at low k permanently
merge and split up over time, which leads to exchange of directional information across the
collective on long time scales, eventually leading to global consensus and high coordination
levels.
This situation changes with increasing attention capacity k. The chance of an agent to
7
0.00.20.40.60.81.0Rinf0.00.20.40.60.81.0abc05101520attention limitk0.00.20.40.60.81.0accuracy CDS0.050.10.150.20.250.00.20.40.60.81.0Rinf0.00.20.40.60.81.0accuracy Ck1361224DS= 0DS= 0.2detect a DS increases strongly, as the distance to the k nearest objects is an increasing
function of k. As soon as this distance becomes larger than the distance to the next DS
(dil (cid:46) 1), agents are capable to detect the environmental cues reliably, and react to them
if they are within their repulsion zone. In addition, larger k also increases the number of
other agents paying attention to a direct-responder. The motion of the emergent sub-groups
is still well coordinated at a local scale (see Supplementary Movie S2). However, these
sub-groups interact now predominantly with the environment by changing their direction of
motion and by complex fission-fusion dynamics directly triggered by the DSs. This results
in quick loss of directional information across time and space, and in a vanishing impact
of the directional information provided by the informed individuals, which yields a strong
decrease in collective accuracy C.
The contribution of both effects to the emergent self-isolation for low k is shown in Fig
4a: The fraction of direct responders rd is much lower for low k and shows only a slow
increase with increasing DS density ρDS. The same holds for the fraction of first-order
indirect responders ri, defined by agents paying attention to at least one direct responder.
For k = 2, ri < 0.1 for all DS densities studied, whereas for k = 24 at high DS densities, it
saturates at more than half of the entire collective (ri ≈ 0.55).
Collective Avoidance of Repulsion Zones. While limited attention is beneficial for
consensus formation in complex environments, it may be very detrimental to the collective
if the environmental cues (DSs) provide reliable informations about environmental dangers
(e.g. predators).
In order to quantify the performance of the collective to respond to environmental cues, we
introduce in the following a DS avoidance measure. First, we compute the average fraction
of agents in "safe" areas, i.e. outside the DS repulsion zone. As a reference, let us estimate
the same quantity in a control numerical experiment with the same number of agents but
who are only interacting exclusively with the environment and not with other agents. Our
DS avoidance measure A is then defined as the ratio between these two quantities (see
Methods for details). This measure A allows us not only to compare the relative collective
performance of the flock at different attention capacities k, but also to compare it with the
performance of solitary agents without any social interactions. For A > 1, social interactions
provide a benefit with respect to solitary individuals. For A < 1, a collective performs on
average worse than solitary agents in avoiding the potentially dangerous areas. Fig. 4b
shows clearly that for small values of k, the collective performs much worse than solitary
individuals in avoiding the repulsion zones. Here, increasing k results in a monotonous
increase in A. However, we observe A > 1 only for sufficiently large k (cid:38) 12. For lower k it
turns out that social interactions are detrimental with respect to DS avoidance. The observed
behavior is directly linked to the emergent local echo-chamber effect at small k. Socially
interacting individuals can only outperform solitary agents in responding to environmental
cues, if sufficient amount of information is able to enter the interaction network and spread
effectively through it.
We can quantify the emergent trade-off between migration accuracy (quantified by C) and
DS avoidance (measured by A) through a global fitness function F (C, A) (see Methods and
Supplementary Information for details). While F depends implicitly on the attention limit k
through A and C, we introduce the parameter δ to quantify the relative benefits of avoidance
versus accuracy. Whereas in safe environments, benefit of avoidance may be negligible
(δ ≈ 0), in environments where local cues provide important information about potential
dangers one can assume δ (cid:29) 1. Fig. 4c shows F as a function of k and δ for collectives
8
FIG. 4. a: The fraction of agents responding to DS directly rd (direct responders, solid lines),
or indirectly via social interaction with direct responders ri (indirect responders, dashed lines),
for k = 2 (blue) and k = 24 (red) versus DS density ρDS. b: Normalized DS avoidance A
versus attention limit k for different DS densities ρDS. A = 1 corresponds to the DS avoidance of
solitary (non-interacting) agents. c: Global fitness versus attention limit k and relative benefits
of DS avoidance δ at ρDS = 0.25. Red (blue) regions correspond to better (worse) performance
of a collective than isolated individuals according to the fitness function used. d and e: Example
snapshots of emergent collective behavior in structured environments with a circular, DS-free path.
For low attention capacity (k = 1, e), individuals ignore the structure of the environment and align
with the preferred direction of migration. At high attention capacity (k = 16, f), the collective
behavior is dominated by the environmental structure and collective migration breaks down. f:
Normalized DS avoidance A in structured environment depicted in d, e versus attention limit k.
A = 1 is the DS avoidance of solitary agents in the same environment. Rinf = 0.1 in all the panels
in this figure.
in random DS fields. We note that F = 0 corresponds to the average fitness expected for
solitary individuals. For low δ, we observe a single maximum of F at low attention capacities
k ≈ 2. For increasing benefits of DS avoidance δ, a second maximum emerges at large k,
while at low k, the socially interacting collectives are on average outperformed by solitary
individuals.
This trade-off between accuracy and avoidance, or "responsiveness", becomes particularly
prominent if we consider structured, inhomogeneous environments containing free paths and
voids in a landscape otherwise filled with high density of DSs, instead of homogeneous,
random environments. At low k the agents completely ignore the environmental structure
and their dynamics is dominated by social interactions with high directional accuracy. At
high k, the situation reverses and the collective dynamics is dominated by the environment,
where agents track the environmental structure, staying preferentially in areas with low DS
9
0.050.100.150.200.25DS0.00.20.40.6k=2k=24rdriabcdfattention limitk051015200.70.80.91.01.10.050.10.150.20.25random environmente01020attention limitk0.60.81.01.21.4DS avoidance Astructured environmentDS avoidance ADS densityDSdensity, while ignoring the preferred direction of migration (see Fig. 4d -- f, see also Supp.
Fig. S3 and Supp. Movie S3, S4).
Model Variations & Generality of Results. So far, we have considered social inter-
actions which do not pay special attention to neighbors responding to DSs. This is motivated
by the idea of social interactions being based on observations of behavior of others but ab-
sence of direct communication about the cause of their particular behavior. In order to test
the robustness of our results, we explored an extension of the basic model, by introduc-
ing active signaling about potential danger by agents interacting with a DS. In this case,
all neighboring agents connected to the signaler put their full attention on the signaling
agent and respond only to it, while ignoring other non-signaling agents. As expected, the
additional signaling improves the collective DS avoidance due to increased saliency of the
corresponding cues within the network. However, the general results - in particular the
coordination and responsiveness trade-off - remain unchanged (see Supp. Fig. S4).
accuracy C by the (normalized) average velocity C = ((cid:80)
Furthermore, the general trade-off between consensus and responsiveness to DSs does
not depend on the presence of informed individuals. For Rinf = 0 and ρDS = 0, our
system reduces to a Vicsek model with topological interactions in homogeneous environ-
ments [27, 28, 36]. With Rinf = 0 and ρDS > 0, i.e. without the bias provided by informed
individuals, the model becomes a topological Vicsek-like model in heterogeneous environ-
ments. All the phenomena discussed above hold also in this case, if we replace the collective
i ui)/N , which quantifies the overall
degree of (orientational) order in the system (see Supp. Information and Supp. Fig. S5). It
is worth stressing that there exist fundamental differences with metric Vicsek-like models in
heterogeneous environments (cf. [33, 34]), where agents are not subject to any cognitive limi-
tation and display exclusively a limited perception capacity. Finally, we note that variations
of the topological interaction mechanism do not affect the reported results. Specifically, we
study the behavior of a model where the k-nearest objects are selected from the Voronoi
neighborhoods (Supp. Fig. S6). This version of the model resembles the spatially balanced
topological flocking algorithm proposed in [41] and represents a better approximation of
visual networks [14]. All this suggests that the fundamental coordination-responsiveness
trade-off discussed here is independent on the specific choice of the social interaction model.
DISCUSSION
Using a generic flocking model we have demonstrated the importance of finite attention
capacity of individuals for collective information processing in complex environments. In
our model, agents dynamically allocate their limited attention to process social and en-
vironmental stimuli, whereby the saliency of different stimuli is governed by their spatial
vicinity. We demonstrated, that contrary to the general intuition, large-scale coordination
and, as a result, the accuracy of collective migration in complex environments is maximized
for strongly limited individual attention capacity. High levels of accuracy for agents which
can pay attention only to few stimuli at a time, are a direct consequence of a self-isolation
from distracting environmental cues through social interactions. On the other hand, the
increased ability of agents to respond collectively to environmental cues for high attentional
capacity leads to a breakdown of collective accuracy for a minority of informed individuals
as the strong information inflow through local distractions overrides the information on the
global preferred direction of motion available only to a minority of informed individuals.
This demonstrates a fundamental trade-off between large-scale coordination and collective
10
accuracy on the one hand, and the dynamical response to local environmental cues in com-
plex environments, on the other hand. We would like to emphasize that our general finding
of weaker connected networks achieving higher global accuracy in complex environments is
diametrically opposed to widely accepted and intuitive knowledge in network science that
more connections lead more effective information exchange and thus higher levels of syn-
chronization (see e.g. [29, 30, 39, 40]). We recover this intuitive result for flocking in empty
environments, which demonstrates how taking into account environmental disturbances may
dramatically change the collective behavior of self-organizing systems.
Our results suggest a specific link between cognitive and sensory capabilities of flocking
animals and the ecological context. For example, for migrating animals, with high fitness
benefits associated with coordination and information sharing on a preferred migration direc-
tion, with no (or very low) fitness costs of ignoring local environmental cues, strongly limited
attention appears to be beneficial. However, if collective response to environmental cues is
highly relevant for individual fitness but global coordination is not, as for example in forag-
ing reef fish [42], then being able to pay attention to many stimuli simultaneously becomes
important. This yields testable hypotheses, on how the attention capability of different
species exhibiting grouping behavior should co-vary with ecological niche, or how individ-
uals within the same species should modulate their attention capabilities across contexts.
Here, we note that a recent analysis of collective behavior in Hemigrammus rhodostomus,
a strongly schooling fish species, suggests that an individual fish appears to pay attention
only to one or two neighbors at a time [21].
Interestingly, being social offers an advantage over solitary behavior with respect to
response to DSs only above a critical attention capacity. This poses some fundamental
questions regarding the co-evolution of social behavior and individual attention capacity,
especially taking into account potential developmental costs of higher attention capacity.
Overall, our results point towards a complex interrelation between pre-existing attention
capability, evolution of grouping behavior and the ecological niche.
We note that in our minimal model we varied only a single dimension of cognitive capa-
bilities, namely the total number of objects an individual can pay attention to. There are
other more complex cognitive processes, which affect the individual processing of a large
number of social and non-social stimuli. For example object recognition and classification,
may enable individuals to dynamically vary the relative saliency of social and non-social
stimuli, which is not considered in this work but could allow individuals to adapt to dif-
ferent behavioral contexts. In general, the strength of the observed trade-off will depend
on model choice and model details. For example, making the agents more likely to detect
danger sites, will make the collectives more responsive to the environmental cues. However,
if the overall attention capacity has an upper bound - as assumed here - it must come at the
cost of decreasing social interactions. Hence, the existence of the general effects discussed
here, in particular the surprising increase in collective accuracy with decreasing attention
capabilities, is not restricted to particular model choice. The qualitative results should hold
as long as the following three conditions are met: 1) the attention capacity is limited, 2) the
salience of social cues has some distance-based component, and 3) the structure of the in-
teraction emerges naturally from spatial self-organization. Overall, this work demonstrates
the fundamental importance of potential constraints in sensory and cognitive abilities of
individuals on emergence and function of collective behavior.
We considered explicitly the case of spatial flocking behavior, however the coordination-
responsiveness trade-off as a general principle should be observable in different collective
11
behaviors. Only recently it was shown that Guinea baboon exhibit stronger response to
known social cues than to novel ones. It was hypothesised that this unexpected behavior
can be linked to the complex social environments in which Guinea baboon groups live, and
the corresponding necessity to filter out irrelevant or distracting information ("social noise")
[43].
Our findings have also implications for the design of interaction networks in artificial
distributed systems, such as robotic swarms that operate in complex environments. Instead
of continuously increasing the sensory and computational abilities of individual agents in
order to cope with the consensus problems in complex environments, it may be promising to
think about constraining the "cognition" of swarming robots. By generating specifically tai-
lored attentional bottlenecks, resulting in emergent self-isolation as observed here, one can
facilitate coordination and exchange of relevant information in complex environments. At-
tentional bottlenecks based on static features, e.g. colors, which can be easily distinguished
from the background, are widely used in swarm robotics [44]. Here, we demonstrated that
dynamical features (as opposed to static ones), like relative distance, or relative speed [18],
could provide effective means of coordination in complex environments, where "filtering"
based on static features is difficult or not feasible.
Last but not least, our work yields potentially interesting implication for social sciences,
where "echo-chambers" have received considerable attention recently. In human social net-
works, this effect is typically linked to homophily and confirmation bias [45, 46]. Our results
show collective self-isolation from conflicting external information as agents moving in the
same direction form tightly interacting social groups which ignore environmental cues. This
can be viewed as an "echo-chamber"-like effect, which emerges naturally even in the absence
of such explicit biases and self-sorting mechanisms. On the one hand, this suggests that these
self-isolation tendencies may be much more prevalent and easier to obtain for agents with
limited attention capacity. On the other hand, our results provide support for the evolution
of proximate, socio-psychological mechanism facilitating the formation of echo chambers,
such as homophily, by demonstrating how an emergent "echo chambers"-like effect strongly
increase intra-group synchronization for a collective in a complex environment.
METHODS
Agent-based Model. We consider a system of N self-propelled agents and NDS danger
sites DSs in a two dimensional domain of size L × L with periodic boundary conditions
(torus). The agent and DS densities are thus ρ = N/L2 and ρDS = NDS/L2. The agents
move with a fixed speed v0 = 0.5 and respond to other agents and DSs by changing their
direction of motion ui = (cos ϕi(t), sin ϕi(t))T . The behavior of each agent is mathematically
described by following stochastic equations of motion, whereby dϕi/dt corresponds to the
turning rate of agent i:
(2)
(3)
(cid:18) cos ϕi(t)
(cid:19)
sin ϕi(t)
sin(ϕj − ϕi) − γpsin(ϕi)
(cid:35)
j∈kNO
sin(αi,l − ϕi) + ηξi(t)
d(cid:126)xi
dt
dϕi
dt
= (cid:126)vi(t) = v0 ui(t) = v0
(cid:34)
(cid:88)
γs
ns
l∈kNO
= (1 − gi(t))
+gi(t)
γl
nl
(cid:88)
12
The first term in the turning response equation 3, is the alignment interaction. A focal
individual aligns with the strength γs = 1 with neighbors j, which are part of its kNO-set.
In addition, informed individuals have a (weak) tendency γp = 0.1 to move in a preferred
direction, here +x = (1, 0)T (γp = 0 for non-informed individuals). The turning away
(repulsion) from the DS l, which are in the kNO-set is given by the third term (γl = 1). Here,
αi,l is the spatial position angle of the focal agent relative to the DS l. Both interactions are
y∈kNO 1). The
function gi(t) determines whether the agent responds to DSs, or whether it aligns with its
neighbors, and, in the case of informed individuals, biases its motion towards the preferred
direction of motion. It is defined as
normalized by the respective number of agents and DSs, respectively (ny =(cid:80)
(cid:40)
gi(t) =
1 for
0 else .
l ∈ kNO and dil < r
(4)
Note that an agent responding to a non-social cue (g(t) = 1), will not interact socially with
other agents until its interaction with the DS is terminated, either because it leaves the
repulsion zone or the site falls out of its k-nearest object set. The last term in equation 3
accounts for the stochasticity in the motion of individuals, with η being the angular noise
strength and ξi(t) a normally distributed Gaussian white noise with standard deviation 1.
In all the simulations discussed, η = 0.25 and the average density of particles is fixed at 1
in a box of linear size L = 25 (N = 625). We have confirmed, that a detailed choice of v0,
η, and L does not change the qualitative observations discussed here.
Avoidance Parameter. We quantify avoidance of repulsion zones through A = 1 −
(cid:104)Nrz(t)/N(cid:105). Here Nrz(t) is the number of agents, which are within at least one repulsion
zone at time t, and (cid:104)·(cid:105) represents temporal average in the stationary state. A will always
decrease with DS density, as more and more space is occupied by repulsion zones. In order
to control for this trivial effect, we rescale A by the corresponding value for non-interacting
agents A = A/ Ani. Thus, A = 1 indicates same average performance of the flock as solitary
agents without any social interactions. Here, solitary individuals are always responding to
a DS, once they are within the corresponding repulsion zone.
Fitness Function. We can quantify the emergent trade-off between collective accuracy
and responsiveness, through the following global fitness function depending on the collective
accuracy C and DS avoidance:
F (C, A) = C + δ(A − 1) ,
(5)
with δ being the relative benefits of DS avoidance with respect to migration accuracy. The
above function was defined in a way so that a value F = 0 corresponds to the behavior of
solitary individuals, where the average accuracy vanishes (C = 0) and the DS avoidance
is A = 1, according to the definition above. Thus, only for F > 0 the collectives perform
better than single individuals (see Supp. Information for more details).
ACKNOWLEDGMENTS
P. Romanczuk & P. Rahmani acknowledge support by the German Science Foundation
(DFG) through RO 4766/2-1 and the German Academic Exchange Service (DAAD). P.
Romanczuk acknowledges funding by the Deutsche Forschungsgemeinschaft (DFG, German
13
Research Foundation) under Germanys Excellence Strategy -- EXC 2002/1 "Science of Intel-
ligence -- project number 390523135. P. Rahmani acknowledges support from the Ministry
of Science, Research and Technology of Iran. F. Peruani was supported by the Agence Na-
tionale de la Recherche via project BactPhys, Grant No. ANR-15-CE30-0002-01. We thank
Ana Sof´ıa Peruani for the artwork in Fig.1a.
[1] Jens Krause, Graeme D Ruxton, and Stefan Krause, "Swarm intelligence in animals and
humans," Trends in Ecology & Evolution 25, 28 -- 34 (2010).
[2] Andreas Huth and Christian Wissel, "The simulation of the movement of fish schools," Journal
of theoretical biology 156, 365 -- 385 (1992).
[3] Tam´as Vicsek, Andr´as Czir´ok, Eshel Ben-Jacob, Inon Cohen, and Ofer Shochet, "Novel type
of phase transition in a system of self-driven particles," Physical review letters 75, 1226 (1995).
[4] Daniel Grunbaum, "Schooling as a strategy for taxis in a noisy environment," Evolutionary
Ecology 12, 503 -- 522 (1998).
[5] Iain D Couzin, Jens Krause, Richard James, Graeme D Ruxton, and Nigel R Franks, "Collec-
tive memory and spatial sorting in animal groups," Journal of theoretical biology 218, 1 -- 11
(2002).
[6] Iain D Couzin, Jens Krause, Nigel R Franks, and Simon A Levin, "Effective leadership and
decision-making in animal groups on the move," Nature 433, 513 (2005).
[7] Felipe Cucker and Cristi´an Huepe, "Flocking with informed agents," Mathematics in Action
1, 1 -- 25 (2008).
[8] Vishwesha Guttal and Iain D Couzin, "Social interactions, information use, and the evolution
of collective migration," Proceedings of the national academy of sciences 107, 16172 -- 16177
(2010).
[9] Andrew Berdahl, Colin J Torney, Christos C Ioannou, Jolyon J Faria, and Iain D Couzin,
"Emergent sensing of complex environments by mobile animal groups," Science 339, 574 -- 576
(2013).
[10] Iain D Couzin, Christos C Ioannou, Guven Demirel, Thilo Gross, Colin J Torney, Andrew
Hartnett, Larissa Conradt, Simon A Levin, and Naomi E Leonard, "Uninformed individuals
promote democratic consensus in animal groups," science 334, 1578 -- 1580 (2011).
[11] Ariana Strandburg-Peshkin, Damien R Farine, Iain D Couzin, and Margaret C Crofoot,
"Shared decision-making drives collective movement in wild baboons," Science 348, 1358 --
1361 (2015).
[12] Mehdi Moussaıd, Dirk Helbing, Simon Garnier, Anders Johansson, Maud Combe, and Guy
Theraulaz, "Experimental study of the behavioural mechanisms underlying self-organization
in human crowds," Proceedings of the Royal Society B: Biological Sciences 276, 2755 -- 2762
(2009).
[13] Mehdi Moussaıd, Dirk Helbing, and Guy Theraulaz, "How simple rules determine pedestrian
behavior and crowd disasters," Proceedings of the National Academy of Sciences 108, 6884 --
6888 (2011).
[14] Ariana Strandburg-Peshkin, Colin R Twomey, Nikolai WF Bode, Albert B Kao, Yael Katz,
Christos C Ioannou, Sara B Rosenthal, Colin J Torney, Hai Shan Wu, Simon A Levin, et al.,
"Visual sensory networks and effective information transfer in animal groups," Current Biology
23, R709 -- R711 (2013).
14
[15] Daniel JG Pearce, Adam M Miller, George Rowlands, and Matthew S Turner, "Role of
projection in the control of bird flocks," Proceedings of the National Academy of Sciences
111, 10422 -- 10426 (2014).
[16] Daniel Strombom, "Collective motion from local attraction," Journal of theoretical biology
283, 145 -- 151 (2011).
[17] Lucas Barberis and Fernando Peruani, "Large-scale patterns in a minimal cognitive flocking
incidental leaders, nematic patterns, and aggregates," Physical review letters 117,
model:
248001 (2016).
[18] Pawel Romanczuk and Lutz Schimansky-Geier, "Swarming and pattern formation due to
selective attraction and repulsion," Interface focus 2, 746 -- 756 (2012).
[19] Maksym Romensky, Vladimir Lobaskin, and Thomas Ihle, "Tricritical points in a vicsek
model of self-propelled particles with bounded confidence," Physical Review E 90, 063315
(2014).
[20] Michele Ballerini, Nicola Cabibbo, Raphael Candelier, Andrea Cavagna, Evaristo Cisbani,
Irene Giardina, Vivien Lecomte, Alberto Orlandi, Giorgio Parisi, Andrea Procaccini, et al.,
"Interaction ruling animal collective behavior depends on topological rather than metric dis-
tance: Evidence from a field study," Proceedings of the national academy of sciences 105,
1232 -- 1237 (2008).
[21] Li Jiang, Luca Giuggioli, Andrea Perna, Ram´on Escobedo, Valentin Lecheval, Cl´ement Sire,
Zhangang Han, and Guy Theraulaz, "Identifying influential neighbors in animal flocking,"
PLoS computational biology 13, e1005822 (2017).
[22] James Intriligator and Patrick Cavanagh, "The spatial resolution of visual attention," Cogni-
tive psychology 43, 171 -- 216 (2001).
[23] J Jay Todd and Ren´e Marois, "Capacity limit of visual short-term memory in human posterior
parietal cortex," Nature 428, 751 (2004).
[24] Patrick Cavanagh and George A Alvarez, "Tracking multiple targets with multifocal atten-
tion," Trends in cognitive sciences 9, 349 -- 354 (2005).
[25] C Shawn Green and Daphne Bavelier, "Enumeration versus multiple object tracking: The
case of action video game players," Cognition 101, 217 -- 245 (2006).
[26] George A Alvarez and Steven L Franconeri, "How many objects can you track?: Evidence for
a resource-limited attentive tracking mechanism," Journal of vision 7, 14 -- 14 (2007).
[27] Francesco Ginelli and Hugues Chat´e, "Relevance of metric-free interactions in flocking phe-
nomena," Physical Review Letters 105, 168103 (2010).
[28] Yen-Liang Chou, Rylan Wolfe, and Thomas Ihle, "Kinetic theory for systems of self-propelled
particles with metric-free interactions," Physical Review E 86, 021120 (2012).
[29] Yilun Shang and Roland Bouffanais, "Influence of the number of topologically interacting
neighbors on swarm dynamics," Scientific reports 4, 4184 (2014).
[30] BH Lemasson, JJ Anderson, and RA Goodwin, "Collective motion in animal groups from
a neurobiological perspective: the adaptive benefits of dynamic sensory loads and selective
attention," Journal of Theoretical Biology 261, 501 -- 510 (2009).
[31] Ashley JW Ward, James E Herbert-Read, David JT Sumpter, and Jens Krause, "Fast and
accurate decisions through collective vigilance in fish shoals," Proceedings of the National
Academy of Sciences 108, 2312 -- 2315 (2011).
[32] Ariana Strandburg-Peshkin, Damien R Farine, Margaret C Crofoot, and Iain D Couzin,
"Habitat and social factors shape individual decisions and emergent group structure during
baboon collective movement," Elife 6, e19505 (2017).
15
[33] Oleksandr Chepizhko, Eduardo G Altmann, and Fernando Peruani, "Optimal noise maxi-
mizes collective motion in heterogeneous media," Physical Review Letters 110, 238101 (2013).
[34] Oleksandr Chepizhko and Fernando Peruani, "Active particles in heterogeneous media display
new physics," The European Physical Journal Special Topics 224, 1287 -- 1302 (2015).
[35] Etienne Sirot, "Social information, antipredatory vigilance and flight in bird flocks," Animal
Behaviour 72, 373 -- 382 (2006).
[36] Anton Peshkov, Sandrine Ngo, Eric Bertin, Hugues Chat´e, and Francesco Ginelli, "Contin-
uous theory of active matter systems with metric-free interactions," Physical review letters
109, 098101 (2012).
[37] Tam´as Vicsek and Anna Zafeiris, "Collective motion," Physics reports 517, 71 -- 140 (2012).
[38] M Cristina Marchetti, Jean-Fran¸cois Joanny, Sriram Ramaswamy, Tanniemola B Liverpool,
Jacques Prost, Madan Rao, and R Aditi Simha, "Hydrodynamics of soft active matter,"
Reviews of Modern Physics 85, 1143 (2013).
[39] Igor V Belykh, Vladimir N Belykh, and Martin Hasler, "Blinking model and synchronization
in small-world networks with a time-varying coupling," Physica D: Nonlinear Phenomena 195,
188 -- 206 (2004).
[40] Florian Dorfler, Michael Chertkov, and Francesco Bullo, "Synchronization in complex os-
cillator networks and smart grids," Proceedings of the National Academy of Sciences 110,
2005 -- 2010 (2013).
[41] Marcelo Camperi, Andrea Cavagna, Irene Giardina, Giorgio Parisi, and Edmondo Silvestri,
"Spatially balanced topological interaction grants optimal cohesion in flocking models," In-
terface focus 2, 715 -- 725 (2012).
[42] Andrew M Hein, Michael A Gil, Colin R Twomey, Iain D Couzin, and Simon A Levin, "Con-
served behavioral circuits govern high-speed decision-making in wild fish shoals," Proceedings
of the National Academy of Sciences 115, 12224 -- 12228 (2018).
[43] Lauriane Faraut and Julia Fischer, "How life in a tolerant society affects the attention to
social information in baboons," Animal Behaviour 152, 11 -- 17 (2019).
[44] Manuele Brambilla, Eliseo Ferrante, Mauro Birattari, and Marco Dorigo, "Swarm robotics:
a review from the swarm engineering perspective," Swarm Intelligence 7, 1 -- 41 (2013).
[45] Pranav Dandekar, Ashish Goel, and David T Lee, "Biased assimilation, homophily, and the
dynamics of polarization," Proceedings of the National Academy of Sciences 110, 5791 -- 5796
(2013).
[46] Michela Del Vicario, Alessandro Bessi, Fabiana Zollo, Fabio Petroni, Antonio Scala, Guido
Caldarelli, H Eugene Stanley, and Walter Quattrociocchi, "The spreading of misinformation
online," Proceedings of the National Academy of Sciences 113, 554 -- 559 (2016)
16
SUPPLEMENTARY INFORMATION
Stationary distribution of in and out-degrees
From our numerical simulations, we can extract the stationary probability distribution
f (Dout, Din) for an agent having a particular combination of out-degree Dout and in-degree
Din.
In homogeneous environments, ρDS → 0, the out-degree of each agent is directly set by
the attention capacity k. Thus, the distribution of individual in and out-degrees f (Dout, Din)
is sharply peaked at Dout = k. The in-degree is not fixed: Whereas a focal individual i pays
attention only to k nearest neighbors, the number of other agents m paying attention to it
may be lower or higher. Therefore, we observe a spreading out of f (Dout, Din) along the in-
degree axis. This pattern holds also for finite number of DSs, however, now in addition, there
is also a finite probability that an agent interacting with a DS ignores its neighbors (see Fig.
2c,d). In this case, its out-degree with respect to social interactions is zero: Dout = 0, while
the in-degree can assume many possible values depending on how many other agents pay
attention to it at a given time. In general, for a finite DS density, the combined distribution
f (Dout, Din) shows a bimodal distribution with two maxima at Dout = 0 and Dout = k.
For low k, the fraction of agents responding to DSs remains low even at very high ρDS.
The average out-degree as well as the average in-degree increases as expected with increasong
k. However, the expected in-degree of direct responders with Dout = 0 is always lower then
the expected in-degree of other agents with Dout > 0. As a results, for low k < 3, DS
responders become often completely isolated from other agents by having Din = Dout = 0.
Emergence of global order in the absence of informed individuals
In the absence of informed individuals Rinf = 0, there is no preferred direction of motion.
Here, instead of collective accuracy, coordination can be quantified using average polarization
of the flock
(cid:43)
ui
,
(6)
(cid:42) N(cid:88)
i=1
C =
1
N
which is equivalent to the ferromagnetic order parameter in physics. By replacing C by C,
we observe the same fundamental coordination-responsiveness trade-off as discussed in the
main text (Fig. S5).
Quantifying the Coordination-Responsiveness Trade-Off -- Global Fitness Function
We can quantify the emergent trade-off between coordination and responsiveness, by
introducing the following global fitness function depending on the collective accuracy C and
DS avoidance A:
F (C, A, k) = bcC + ba(A − 1)
(7)
with bc being the benefits per unit of directional consensus, and ba the benefits per unit of
DS avoidance. Without loss of generality, the equation can be rescaled, by bc to
F (C, A) = C + δ(A − 1),
(8)
17
where δ = ba/bc represents the relative benefits of DS avoidance versus coordination. Please
note that F is not fitness in a strict evolutionary sense (selection at the level of individuals),
as it is a function collective variables A and C. Here, "fitness" refers rather to a collective
utility function, where (local) maxima correspond to (local) optimal collective strategies
for homogeneous collectives. For a given set of parameters, A and C are calculated in a
stationary regime of a stochastic system. Thus both variables reflect the average ability of
individuals to avoid DSs and to coordinate with their neighbors in homogeneous groups.
For δ = 1 both behaviors yield equal benefits, whereas for δ > 1 (δ < 1) DS avoidance
is more (less) beneficial than coordination. For a wide range of DS densities and angular
noise, we observe two distinct maxima of F in the δ, k-plane. For δ (cid:28) 1 coordination is
much more beneficial than DS avoidance and the maximum of F is located at small k where
coordination is highest. With increasing δ this maximum decreases and we eventually see
a rise of a second maximum at large k, leading to two local maxima at intermediate δ.
Eventually, for large δ, where DS avoidance becomes far more important than coordination,
we observe a single global maximum at large k.
Numerical Implementation and Experiments
The mathematical model was implemented in C/C++ with the k nearest object inter-
action implemented using the kd-tree implementation in the CGAL library [1]. The kNN
interaction implementation in periodic boundary conditions was combining the standard
kd-tree based algorithm with generation of mirror images of the environment to account for
the interactions on a torus.
The stochastic equations of motion were numerically integrated using a standard Euler-
Maruyama scheme [2]. The numerical time step was set dt = 0.1, whereby also smaller time
steps were tested to show that the general results do not depend on the time step.
All the data were obtained by averaging over 20 realizations, each including 5 × 104
relaxation time steps and 105 stationary time steps.
[1] M. Basken. 2D Range and Neighbor Search. In CGAL User and Reference Manual. C GAL
Editorial Board, 4.13 edition, 2018
[2] R. Mannella. Integration of stochastic differential equations on a computer. I nternational
Journal of Modern Physics C 13(09) 1177-1194, 2002.
18
SUPPLEMENTARY MOVIES
Each supplementary movie shows the spatial dynamics in the main panel (left), the
corresponding accuracy C versus time (right top) and the corresponding DS avoidance A
versus time (bottom right). The black line at A = 1 corresponds to the DS avoidance
of non-interacting agents (see Materials and Methods). The first part shows the initial
development of the system (t = 0− 200), the second part shows the stationary state at large
times (t = 5200 − 5400).
Supplementary Movie 1: Collective behavior at high density of DSs for low attention
capacity k = 1 characterized by high accuracy of collective migration.
Supplementary Movie 2: Collective behavior at high DS densities for high attention
capacity k = 24 characterized by efficient response to environmental cues.
Supplementary Movie 3: Collective behavior in structured environment with a circular
DS free region for low attention capacity k = 1.
Supplementary Movie 4: Collective behavior in structured environment with a circular
DS free region for high attention capacity k = 24.
SUPPLEMENTARY FIGURES
19
FIG. S1. Snapshots of the (undirected) social interaction network in random environments with
ρDS = 0.25 for k = 3 upper panel, and k = 12 lower panel, at Rinf = 0.1. Black agents are socially
interacting, and red agents react to DSs. Informed and uninformed individuals are represented by
empty and filled circles, respectively. Light blue circles are repulsion zones of DSs specified with
blue dots. For the sake of clarity, the links between agents interacting with their periodic neighbors
are removed. The black squares depict the close-ups shown in panels a, b of Fig. 2 (main text).
For low k(= 3) the network is sparse, composed of many small connected components, whereas for
large k = 12, the network is highly connected with less components.
20
FIG. S2. Temporal interaction networks: Average number of connected components of the inter-
action network versus attention limit k (a). For all DS densities ρDS, we observe a fast decay of
the number of connected components, which due to constant number of agents N is equivalent to
the growth of the average connected component size, indicating a more tightly connected temporal
network. The average life time of an edge in the interaction network decreases with increasing
density of DSs (b). However with increasing k for a fixed DS density we observe longer life times
due to increased connectivity in the interaction network.
21
0.050.100.150.200.25DS345678average edge life timek124816ab0510152025attention limitk050100150200250average number of CC0.050.100.150.200.25DSFIG. S3. Collective behavior of agents (Rinf = 0.1) in a structured environment with circular DS
free path. A: Accuracy C (triangles) and normalized DS avoidance A (circles) versus attention
limit k. The horizontal line, A = 1, corresponds to DS avoidance of non-interacting agents. For
socially interacting agents with low k values (k = 1, 2), we observe high accuracy C together with
almost complete ignorance towards environmental cues. By increasing k, more agents start to sense
the environment and react to DSs. At high k, the collective behavior is fully determined by the
local environmental features: We observe collective rotation along the circular path and complete
ignorance of the global migration direction accessible to informed individuals (see Supp. Movie S4).
This trade-off is shown quantitatively by the global fitness function in panel B versus attention
capacity k and relative DS avoidance benefit δ. There are two maxima in global fitness, one for
low k, δ (cid:28) 1, showing migration accuracy to be beneficial for the group, the other at high k, and
δ > 1, which indicates higher benefits associated with DS avoidance in comparison to collective
accuracy.
FIG. S4. Attention trade-off in a group of agents with active signalers. Each agent connected to
another individual signalling direct interaction with a DS (direct responder), pays only attention
to the signaller(s) and ignores other social cues. Accuracy C (a) and normalized DS avoidance A
(b) versus attention limit k for different DS densities at Rinf = 0.1.
22
abglobal fitness05101520attention limitk0.00.20.40.60.81.0accuracy CC0.60.81.01.21.4ADS avoidance A05101520attention limitk0.00.20.40.60.81.0accuracy C0.050.10.150.20.25ab05101520attention limitk0.70.80.91.01.1DS avoidance ADSFIG. S5. Emergence of global order in the system with no informed individuals, Rinf = 0. a: Co-
ordination C (directional order) versus attention limit k for different DS densities. b: Normalized
DS avoidance versus attention limit k. A = 1 corresponds to non-interacting agents. The quali-
tative behavior with a coordination-responsiveness trade-off is similar to the model with informed
individuals, but here instead of a specific direction, the emergent consensus direction is a random
(spontaneous symmetry breaking).
FIG. S6. Collective motion of agents with Voronoi-based kNN interaction network. Here, k nearest
agents are selected from first shell of Voronoi neighbors. If the number of neighbors in first layer
is smaller than k, then depending on k, the second Voronoi shell is considered. It is defined by the
Voronoi neighborhood of the (direct) Voronoi neighbors of the focal agent. Accuracy C (a) and
normalized DS avoidance A (b) versus attention limit k at Rinf = 0.1.
23
05101520attention limitk0.00.20.40.60.81.0coordinationCab05101520attention limitk0.70.80.91.01.1DS avoidance A0.050.10.150.20.25DS05101520attention limitk0.00.20.40.60.81.0accuracy Cab05101520attention limitk0.70.80.91.01.1DS avoidance A0.050.150.25DS |
1803.02797 | 1 | 1803 | 2018-03-07T18:11:30 | Non-equilibrium scaling behaviour in driven soft biological assemblies | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech"
] | Measuring and quantifying non-equilibrium dynamics in active biological systems is a major challenge, because of their intrinsic stochastic nature and the limited number of variables accessible in any real experiment. We investigate what non-equilibrium information can be extracted from non-invasive measurements using a stochastic model of soft elastic networks with a heterogeneous distribution of activities, representing enzymatic force generation. In particular, we use this model to study how the non-equilibrium activity, detected by tracking two probes in the network, scales as a function of the distance between the probes. We quantify the non-equilibrium dynamics through the cycling frequencies, a simple measure of circulating currents in the phase space of the probes. We find that these cycling frequencies exhibit power-law scaling behavior with the distance between probes. In addition, we show that this scaling behavior governs the entropy production rate that can be recovered from the two traced probes. Our results provide insight in to how internal enzymatic driving generates non-equilibrium dynamics on different scales in soft biological assemblies. | physics.bio-ph | physics |
Non-equilibrium scaling behaviour in driven soft biological assemblies
Federica Mura,1, ∗ Grzegorz Gradziuk,1, ∗ and Chase P. Broedersz1, †
1Arnold-Sommerfeld-Center for Theoretical Physics and Center for NanoScience,
Ludwig-Maximilians-Universitat Munchen, D-80333 Munchen, Germany.
(Dated: March 8, 2018)
Measuring and quantifying non-equilibrium dynamics in active biological systems is a major chal-
lenge, because of their intrinsic stochastic nature and the limited number of variables accessible
in any real experiment. We investigate what non-equilibrium information can be extracted from
non-invasive measurements using a stochastic model of soft elastic networks with a heterogeneous
distribution of activities, representing enzymatic force generation. In particular, we use this model
to study how the non-equilibrium activity, detected by tracking two probes in the network, scales as
a function of the distance between the probes. We quantify the non-equilibrium dynamics through
the cycling frequencies, a simple measure of circulating currents in the phase space of the probes.
We find that these cycling frequencies exhibit power-law scaling behavior with the distance between
probes. In addition, we show that this scaling behavior governs the entropy production rate that can
be recovered from the two traced probes. Our results provide insight in to how internal enzymatic
driving generates non-equilibrium dynamics on different scales in soft biological assemblies.
Cells and tissue constitute a class of non-equilibrium
many-body systems [1–5].
Indeed, non-equilibrium ac-
tivity has been observed in various biological systems,
including membranes [6, 7], chromosomes [8], and the cy-
toplasm [9–11]. A distinguishing physical feature of such
biological assemblies is that they are driven out of equi-
librium collectively by internal enzymatic processes that
break detailed balance at the molecular scale. The active
nature of living matter on larger scales can be determined
non-invasively by observing the steady-state stochastic
dynamics of mescoscopic degrees of freedom using time-
lapse microscopy experiments: The non-equilibrium dy-
namics of these systems can manifest as circulating prob-
ability currents in a phase space of mesoscopic coordi-
nates [2, 12, 14, 15]. However, it remains unclear how
such non-equilibrium measures depend on the spatial
scale on which the measurement is performed. This is-
sue is not only of practical relevance in an experiment, it
is also of fundamental importance: a theoretical under-
standing of the spatial scaling behavior of broken detailed
balance in internally driven systems may reveal how to
extract quantitative information from measurable phase
space currents about the active nature of the system.
Here we consider a simple, yet general model for an in-
ternally driven elastic assembly to study non-equilibrium
scaling behavior. This assembly is driven out of equi-
librium by heterogeneously distributed stochastic forces,
representing internal enzymatic activity (Fig. 1). We
quantify the non-equilibrium dynamics of such an assem-
bly by the cycling frequencies associated to steady-state
circulating currents in phase space [14, 15]. To study
how broken detailed balance manifests on different scales
in a given system, we investigate how the cycling fre-
quency of a pair of tracer probes depends on the spatial
distance between these probes. Interestingly, the cycling
frequencies in our model exhibit a power-law scaling with
the distance between probes with an exponent that de-
FIG. 1. Schematic illustrating soft viscoelastic networks with
heterogeneous driving for various types of cellular systems.
A) chromosome B) red blood cell membrane C) cytoskeletal
network with in D-F associated bead-spring models with het-
erogeneous active driving. The color of the bead indicates
the intensity of activity, representing the variance (increasing
from blue to red) of the associated active noise process.
pends on the dimensionality of the system. To provide
a conceptual understanding of this scaling behavior, we
develop an analytical calculation of these exponents. Fur-
thermore, we show that the exponent associated to the
power law of the cycling frequencies also underlies the
scaling behavior of the entropy production rate that can
be recovered from measured trajectories. Therefore, we
provide a framework to study the spatial scaling behavior
of non-equilibrium measures in soft elastic assemblies.
Our model consists of a d-dimensional elastic network
immersed in a simple Newtonian liquid
of N beads,
at temperature T [16–18]. We assume a lattice struc-
ture where each bead is connected to its nearest neigh-
bours by springs of elastic constant k, as illustrated in
Fig. 1. For simplicity, we model internal enzymatic ac-
tivity by a Gaussian white noise with variance αi at
bead i. By assuming white noise, we effectively consider
the dynamics of biological systems on time scales much
longer than the characteristic timescales of the active
processes [14, 19, 20].
Importantly, these activity am-
plitudes, αi ≥ 0, are spatially heterogeneous, reflecting
a spatial distribution of active processes in the system.
These activity amplitudes are drawn independently from
a distribution pα with mean ¯α < ∞ and standard devi-
ation σα < ∞ for each realization of the system. This
description of a heterogeneously driven assembly is simi-
lar to bead-spring models in which the beads are coupled
to distinct heat baths at different temperatures [21–23].
The temporal evolution of the probability distribution,
p(x, t), of the beads' displacements x, relative to their
rest positions, is governed by a Fokker-Planck equation:
∂p(x, t)
∂t
= −∇ · [Axp(x, t)] + ∇ · D∇p(x, t),
= −∇ · j(x, t)
(1)
kB (T +αi)
γ
where j(x, t) = Axp(x, t) + D∇p(x, t) is the probability
current. Here, A is the elastic interaction matrix, incor-
porating all nearest neighbor spring interactions between
beads; the mobility matrix is assumed to be diagonal to
exclude hydrodynamic interactions between the beads,
and is absorbed in A. The diffusion matrix, D, is di-
agonal with elements dij = δij
, where γ is the
damping coefficient describing the viscous interaction be-
tween a bead and the immersing liquid. The steady-state
dynamics of this active network is described by
2 xT C−1x,
(2)
where C = (cid:104)x ⊗ x(cid:105) is the covariance matrix, which
can be obtained by solving the Lyapunov equation
AC + CAT = − 2D [24].
In the simplest limit,
the activities are spatially homogeneous: αi = α ∀ i, re-
sulting in a dynamics that reduce to that of an effective
equilibrium scenario with j = 0. By contrast, in het-
erogeneously driven systems with non-identical αi's, we
obtain Non-Equilibrium Steady-State (NESS) dynamics
with j (cid:54)= 0 [24].
(cid:112)(2π)dN det C
p(x) =
1
e− 1
If we were able to observe the stochastic motion of all
beads in the network, we could directly measure the full
probability current j(x) and extract information about
the complete non-equilibrium dynamics of the system.
However, in an actual experiment typically only a small
subset of the degrees of freedom can be tracked (Fig. 2A).
What information on the non-equilibrium dynamics of
the system can be extracted from such limited observa-
tions? To address this question, we investigate a scenario
where only a few degrees of freedom are accessible.
2
FIG. 2. Reduced system of tracked probed. A) Schematic of
two fluorescently labelled probe beads in a larger system. B)
Elastic force acting on bead j obtained at different time steps
of a simulation of the Langevin dynamics of the full system
(blue points), and the effective linear force, Aeffxr, from ana-
lytical calculations (light blue plane). C) Probability density
(color map) and probability current (white arrows) calculated
analytically from the effective 2D system, together with re-
sults from simulating the full system in the inset. D) The
non-conservative part of the effective force field: (Aeff−AT
xr
(black arrows) can contribute to the rotation in phase space
in non-equilibrium systems.
eff)
2
distribution, pr(xr) = (cid:82) dxk(cid:54)∈[r]p(x1, x2, .., xdN ), of a
We start by reducing our description to the marginal
subset [r] of n tracked degrees of freedom xr. By in-
tegrating out the subset [l] of m unobserved degrees of
freedom xl on both sides of Eq. (1) in the steady-state
limit, we obtain (see supplementary material):
0 = −∇ · [Aeffxrpr(xr)] + ∇ · D[r,r]∇pr(xr),
(3)
where the sub-index [r, r] of a matrix indicates the sub-
matrix corresponding to the reduced set of observed vari-
ables. In addition, we introduce the effective linear in-
teraction (Fig. 2B), which can be written as Aeffxr, with
Aeff = A[r,r] + A[r,l]C[l,r]C−1
[r,r]. Here, A[r,l] and C[l,r] are
rectangular matrices of sizes [n×m] and [m×n], given by
the elements of indices [r, l] of A and [l, r] of C, respec-
tively. Thus, we obtain an effective stationary Fokker-
Planck equation for the reduced system (Eq. (3)). By
solving this equation, we obtain the exact steady-state
reduced probability distribution pr(xr) and probability
current density:
jr(xr) = Aeffxrpr(xr) + D[r,r]C−1
[r,r]xrpr(xr),
(4)
which can, in principle, be measured directly from the
trajectories of the observed degrees of freedom at steady
state (Fig. 2C).
We can use this reduced description to investigate how
broken detailed balance manifests at different scales in
the network. In particular, we consider the simplest case
of a reduced system of only two tracked beads in a larger
system, as illustrated in Fig. 2A. It is convenient to quan-
tify the probability currents in the 2D phase space of
these two tracer beads by a pseudoscalar quantity: the
average cycling frequency around the origin [13–15]. For
linear systems, we can express the reduced probability
current as jr(xr) = Ωrxrpr(xr), where Ωr is a 2D ma-
trix with purely imaginary eigenvalues λ = ±iω, with ω
representing the cycling frequency.
3
(cid:112)(cid:104)ω2(r)(cid:105)α for pairs of beads separated by a distance r.
This cycling frequency can be measured for a pair of
probe beads at a distance r, and this frequency will de-
pend on the specific configuration of the activity ampli-
tudes αi at all beads in the system. We aim to compute
how this cycling frequency depends on r after averag-
ing over all activity configurations. Since ω is expected
to be distributed symmetrically around 0, we calculate
Here, the average (cid:104)...(cid:105)α is taken over an ensemble of activ-
ities {αi} drawn from the distribution pα. Intuitively, the
magnitude of the circulation of currents in phase space
typically decreases with the distance between the probes,
as shown in Fig. 3A. This reduction of the circulation is
reflected by a decrease of the cycling frequency ω with
distance. Remarkably,(cid:112)(cid:104)ω2(r)(cid:105)α appears to depend on
(cid:112)(cid:104)ω2(r)(cid:105)α ∝ r−µ, with µ ≈ 1.9 for a 1D chain with a
the distance between the tracer beads, r, as a power law,
folded Gaussian or an exponential distribution of activi-
ties, as depicted in Fig. 3B.
To investigate how the architecture of the system af-
fects the scaling behavior of the cycling frequencies, we
considered different network structures, including square,
In particular, we calcu-
triangular, and cubic lattices.
lated (cid:112)(cid:104)ω2(r)(cid:105)α, where we also averaged over different
lattice directions.
Interestingly, we find that the char-
acteristic exponent µ appears to depend strongly on the
dimensionality of the lattice, but not on its geometry,
as shown in Fig. 3B-C. These results suggest that the
distance dependence of the cycling frequency is deter-
mined in part by the long wavelength elastic properties
of the system. Importantly, however, the scaling of cy-
cling frequency is sensitive to the spatial structure of the
activities. For example, in the simple case of a delta-
distributed (single-source) activity on a 1D chain, we find
µsingle ≈ 2.4 (Fig. 3B) in contrast to the value 1.9 ob-
tained above for spatially distributed activities.
To obtain more insight into the scaling behavior of the
cycling frequencies, we derive an analytical expression
for the cycling frequency as a function of the distance
between the observed beads, ω(r).
In general, it can
be shown that for a linear system described by a Fokker-
Planck equation, the cycling frequencies are given by (see
supplementary materials):
(cid:104)τij(cid:105)(cid:112)det C[r,r]
ωij =
1
2γ
(5)
FIG. 3. Spatial scaling behavior of cycling frequencies. A)
Steady-state current cycles in phase space of two tracer
beads for a nearby pair of probes (left) and distant pair of
probes (right). B) Scaling behavior of the cycling frequencies,
(cid:112)(cid:104)ω2(r)(cid:105), of a pair of probes beads as a function of their spa-
of the cycling frequencies, (cid:112)(cid:104)ω2(r)(cid:105), obtained for different
tial distances, obtained for a 1D chain and different activity
distributions, as indicated in the legend. C) Scaling behavior
lattices and a folded Gaussian activity distribution. Triangu-
lar and square markers represent triangular and square/cubic
lattices, respectively. Light/dark blue triangles represent tri-
angular networks with zero/finite rest length springs.
where τij := x×fr(x) = xifj(x)−xjfi(x) is a generalized
phase space torque in the xi-xj plane, with fi(x) denot-
ing the deterministic force acting on the ith bead. This
result is intuitive: for an overdamped system the mean
angular velocity is proportional to the mean torque and
the factor 1/(cid:112)det C[r,r] ensures coordinate invariance.
For the 1D chain of beads (Fig. 1D), Eq. (5) reduces to:
(cid:101)∂2
(cid:112)det C[r,r]
2 cij
ωij =
k
γ
,
(6)
denoted as: (cid:101)∂2
where cij is the i, jth element of the covariance matrix
C, and with the discrete second derivative across rows
2 cij = ci,j+1 − 2ci,j + ci,j−1. Thereby, we
have reduced the problem of calculating ω(r) to finding
the covariance matrix of the system.
The structure of D suggests a natural decomposition
of the covariance matrix C into equilibrium (C) and
non-equilibrium (C∗) contributions: C = (kBT /k)C +
(kB ¯α/k)C∗, such that C and C∗ are dimensionless. Both
C and C∗ can be found by solving the Lyapunov equa-
tion, which for the 1D chain is given by
1 cij +(cid:101)∂2
(cid:101)∂2
(cid:101)∂2
ij +(cid:101)∂2
1 c∗
2 cij = −2δij
ij = −2δij
2 c∗
αi
¯α
,
(7)
(8)
where (cid:101)∂2
1 indicates the discrete second derivative across
columns. These equations represent discrete stationary
diffusion equations, with sources of divergence given by
δij and δij(αi/¯α), respectively. This result prescribes
how a spatial distribution of activities structures the co-
variance matrix.
4
We can make further progress by noting that the prin-
ciple of detailed balance imposes ωij = 0 at thermal equi-
2 cij = 0.
2 c∗
ij,
and then expand this equation up to linear order in ¯α/T
to obtain
librium, which together with Eq. (6) implies (cid:101)∂2
We can, therefore, substitute (cid:101)∂2
2 cij in Eq.(6) by (cid:101)∂2
(cid:101)∂2
(cid:113)
ωij =
(9)
.
ij
2 c∗
det C[r,r]
k
γ
¯α
T
We proceed by calculating C∗ for a given distribution
of activities {αi}. Because of the linearity of Eq. (8),
C∗ is a superposition of steady-state solutions to single-
source problem, i.e. a delta-distribution for which all
but one of the activities would be set to zero. Denoting
the element of C∗ at a distance r from the single activ-
ity source by c∗(r), we obtain the "covariance current"
∂rc∗(r) ∼ 1/r. Here we employed a continuous approxi-
mation of the discrete diffusion problem in Eqs. (7) and
(8). Thus, c∗(r) = −a ln(r)+b for a single-source problem
with integration constants a and b. Using this expression
for c∗(r) together with Eq. (9), we obtain for the single
source case: ω2
, where α is
the source's activity.
Next, we use a superposition of single source solutions
for c∗(r) to obtain the non-equilibrium contribution of
the covariance matrix C∗ for a specific configuration of
many activity sources {αi}. Using this result in conjunc-
tion with Eq. (9) and performing an ensemble average
over the distribution of activity realizations, we arrive at
the central result
single(r) = k2
det C[r,r](r)
α2
T 2
a2
r4
γ2
1
(cid:104)ω2(r)(cid:105)α =
k2
γ2
σ2
α
T 2
πa2
2r3
1
det C[r,r](r)
.
(10)
Finally, we note that the elements of the equilibrium co-
variance matrix are given by ci,j = min(i, j)−ij/(N +1),
and find that for r (cid:28) N , det C[r,r](r) exhibits a power law
behavior, det C[r,r](r) ∼ r. Therefore, from this analysis
we find for a 1D chain with heterogenous activities µ = 2,
independent of the activity distribution pα. Furthermore,
we find µsingle = 2.5 for a single-source activity, in accord
with our numerical result (see Fig.3B). This calculation
provides insight into how a combination of features of
the equilibrium and non-equilibrium contributions to the
covariance matrix determine the spatial scaling behavior
of cycling frequencies.
Non-zero cycling frequencies directly reflect broken de-
tailed balance, suggesting a connection between ω and
measures of the internal driving, including the rate of
r
FIG. 4. Spatial scaling behavior of the entropy production
rate, Π2D
, of a pair of probe beads as a function of their
spatial distance r, obtained for different lattices and a folded
Gaussian activity distribution. Note the entropy production
rate of the reduced system is scaled by the total entropy
production rate of the whole network, Πtot. Triangular and
square markers represent triangular and square/cubic lattices,
respectively. Light/dark blue triangles represent triangular
networks with zero/finite rest length springs.
entropy production. For a Markovian system described
by a Fokker-Planck equation, the total entropy produc-
tion rate under steady-state conditions is given by [25]:
Πtot = kB
jT (x)D−1j(x)
p(x)
dx
,
(11)
(cid:90)
where kB is Boltzmann's constant. The validity of this re-
sult relies on the equivalence between the Fokker-Planck
and Langevin descriptions. However, we have seen that
the marginal probability density of the reduced system is
only described by a Fokker-Planck equation (see Eq. (3))
at steady state, reflecting the loss of Markovianity af-
ter coarse-graining. Nonetheless, we can define an effec-
tive dynamics of the reduced set of variables through the
Langevin equation
dxr(t)
dt
= Aeffxr(t) +
2D[r,r] ξr(t),
(12)
with Gaussian white noise ξr(t). This equation of mo-
tion results in the exact steady-state probability and cur-
rent densities, but with an approximate stochastic dy-
namics.
In particular, the effective interaction matrix
Aeff (see Eq. (3)) captures only the average interaction
between the traced variables, as illustrated in Fig. 2B.
Furthermore, in contrast to the full deterministic forces
(Ax), these effective interactions (Fig. 2C) need not to
derive from a potential and, thus, may contain a non-
conservative component (Fig. 2D).
The entropy production rate associated with the effec-
tive Markovian dynamics in Eq.(12) is given by
(cid:113)
(cid:90)
Πr = kB
r (xr)D−1
jT
pr(xr)
[r,r]jr(xr)
dxr
≤ Πtot,
(13)
r = kBω2 Tr (C[r,r]D−1
Π2D
[r,r]).
(14)
[9] A. W. C. Lau, B. D. Hoffman, A. Davies, J. C. Crocker,
5
[5] C. F. Schmidt, and F. C. MacKintosh Curr. Opin. Cell
Biol. 22, 29 (2010).
[6] T. Betz, M. Lenz, J. F. Joanny, C. Sykes, Proc. Natl.
Acad. Sci. 36, 106 (2009).
[7] H. Turlier, D. A. Fedosov, B. Audoly, T. Auth, N. S. Gov,
J. -F. Joanny, G. Gompper, T. Betz Nature Physics 12,
513-519 (2016).
[8] S. C. Weber, A. J. Spakowitz, J. A. Theriot, Proc. Natl.
Acad. Sci. 19, 109 (2010).
T. C. Lubensky Phys. Rev. Lett. 19, 91 (2003).
[10] M. Guo, A. J. Ehrlicher, M. H. Jensen, M. Renz, J. R.
Moore, R. D. Goldman, J. L. Schwartz, F. C. Mackintosh,
D. A. Weitz Cell 4, 158 (2014).
[11] N. Fakhri, A. D. Wessel, C. Willms, M. Pasquali, D. R.
Klopfenstein, F. C. Mackintosh, C. F. Schmidt Science
6187, 344 (2014).
[12] C. Battle, C. P. Broedersz, N. Fakhri, J. Howard, C. F.
Schmidt, F. C. MacKintosh Science 6285, 352 (2016).
[13] J. B. Weiss, Tellus A 3, 55 (2003).
[14] J. Gladrow, N. Fakhri, F. C. MacKintosh, C. F. Schmidt,
C. P. Broedersz, Phys. Rev. Lett. 24, 116 (2016).
[15] J. Gladrow, C. P. Broedersz, C. F. Schmidt Phys. Rev.
E 2, 96 (2017).
[16] M. G. Yucht, M. Sheinman, C. P. Broedersz, Soft Matter,
9 (29) (2013).
[17] C. P. Broedersz, F. C. MacKintosh Rev. Mod. Phys, 86
(3) (2014).
[18] X. Mao, T. C. Lubensky Annu. Rev. Condens. Matter
Phys, 9 (0) (2018).
[19] F. C. MacKintosh, A. J. Levine Phys. Rev. Lett, 100 (1)
018104 (2008).
[20] J. Ruostekoski, J. R. Anglin Phys. Rev. Lett, 91 (19)
190402 (2003).
[21] Z. Rieder, J. L. Lebowitz, E.Lieb J. Math. Phys 8 1073
(1967).
[22] F. Bonetto, J. L. Lebowitz, J. Lukkarinen J. Stat. Phys
116 (2004).
[23] G. Falasco, M. Baiesi, L. Molinaro, L. Conti, F. Baldovin
Phys. Rev. E 92 022129 (2015).
[24] H. Risken, Springer, Berlin, Heidelberg, 1996.
[25] U. Seifert, Reports Prog. Phys. 12, 75 (2012).
[26] G. Bisker, M. Polettini, T. R. Gingrich, J. M. Horowitz
J. Stat. Mech. Theory Exp 2017 093210 (2017).
[27] D. Mizuno, C. Tardin, C. F. Schmidt, F. C. MacKintosh
Science 5810, 315 (2007).
[28] O. Lieleg, M. M. A. E. Claessens, A. R. Bauch Soft
Matter, 6 (2) (2010).
[29] M. H. Jensen, E. J. Morris, D. A. Weitz Biochim. Bio-
phys, 1853 (11 0 0 ) (2015).
[30] G. H. Koenderink, Z. Dogic, F. Nakamura, P. M. Bendix,
F. C. MacKintosh, j. H. Hartwig, T. P. Stossel, and D.
A. Weitz, Proc. Natl. Acad. Sci., 106, 36. (2009)
[31] V. Schaller, C. Weber, C. Semmrich, E. Frey, and A. R.
Bausch, Nature, 467, 7311 (2010)
[32] J. Palacci, S. Sacanna, A. P. Steinberg, D. J. Pine, P. M.
Chaikin, Science, 1230020 (2013).
[33] I. Buttinoni, J. Bialk´e, F. Kummel, H. Lowen, C.
Bechinger, T. Speck, Phys. Rev. Lett., 110(23), p.238301
(2013).
[34] Y. L. Tong, Springer Series in Statistics, Springer-Verlag,
1990.
where jr(xr) is defined in Eq. (4). Note, estimating Πr by
using the Markovian formalism allows us to set a lower
bound for the total entropy production rate Πtot (see sup-
plementary materials), similar to what already shown for
discrete systems [26]. In the 2D case with two traced de-
grees of freedom that we consider here, Eq. (13) reduces
to (see supplementary materials)
This result provides an explicit relation between the par-
tial entropy production rate and the cycling frequency
ω. Note, all quantities in the expression for Π2D
can
be observed in an experiment, providing a direct way
to non-invasively determine the reduced rate of entropy
production for a set of traced degrees of freedom. Since
Tr (C[r,r]D−1
[r,r]) depends only weakly on r, as long as
1 (cid:28) r (cid:28) N , we expect a scaling behavior (cid:104)Π2D
r (cid:105) ∼ r−2µ.
This result shows that the spatial scaling behavior of the
cycling frequencies directly determines the spatial scaling
behavior of the entropy production rate.
r
In summary, we here demonstrate theoretically how
experimental measures of non-equilibrium activity in in-
ternally driven linear networks are affected by the length-
scale at which the system is observed. Specifically, we
developed a general framework to predict the scaling be-
havior of cycling frequencies and the entropy production
rate that can be inferred by tracing pairs of degrees of
freedom. We showed the exponent µ that governs this
behavior for a system with heterogeneous random activ-
ities, is insensitive to the details of distribution of activ-
ities. However, this exponent depends sensitively on the
dimensionality of the system. The predicted scaling be-
haviour can be tested in biological [6, 9–11, 27–31] and
artifical [32, 33] systems under non-equilibrium steady-
state conditions.
We thank E. Frey, J. Gladrow, F. Gnesotto, P. Ron-
ceray, and C. Schmidt for many stimulating discussions.
This work was supported by the German Excellence
Initiative via the program NanoSystems Initiative Mu-
nich (NIM), by a DFG Fellowship through the Gradu-
ate School of Quantitative Biosciences Munich (QBM).
Part of this work was performed at the Aspen Center for
Physics, which is supported by National Science Founda-
tion grant PHY-1607611.
∗ These authors contributed equally
† [email protected]
[1] Fodor E., Marchetti M. C. ArXiv 1708 08652, (2017).
[2] F. Gnesotto, F. Mura, J. Gladrow, C. P. Broedersz Rep.
Prog. Phys. (in press)
[3] F. Julicher, K. Kruse, J. Prost, J.-F. Joanny Physics
Reports 449 1, (2007).
[4] D. Needleman, Z. Dogic Nat. Rev. Mater 2, (2017).
Appendices
DERIVATION OF EQ. (3)
6
Here, we derive Eq. (3), which describes the steady state distribution of traced variables. Integrating out the unob-
served degrees of freedom on both sides of the Fokker-Plank equation (Eq. (1)), and using the Einstein notation for
summing over repeated indexes, we obtain:
(cid:90)
(cid:90)
(I)
dxl∂tp(x) = −
(II)
dxl∂i[aijxjp(x, t)] +
(III)
dxldij∂i∂jp(x, t)
(15)
where aij and dij are the elements of the interaction matrix A and the diffusion matrix D, respectively. Rewriting
the probability as p(x, t) = p(xlxr, t)pr(xr, t), we can separately calculate each term in Eq.(15). The first term (I)
gives:
(cid:90)
(cid:90)
For the second term (II), we obtain
dxl∂tpr(xr, t)p(xlxr, t) = ∂tpr(xr, t)
dxlp(xlxr, t) = ∂tpr(xr, t)
dxl∂i[pr(xr, t)p(xlxr, t)aijxj] = δi,[r]∂i[pr(xr, t)
= δi,[r]∂i[pr(xr, t)aij (cid:104)xjxr, t(cid:105)]
(cid:90)
dxlp(xlxr, t)aijxj]
where δi,[r] = 1 if xi is one of the observed coordinates and zero otherwise. In the first line we use that the probability
density vanishes at infinity faster than 1/x. Similarly, the third term (III) can be written as
dxldij∂i∂j[pr(xr, t)p(xlxr, t)] = δi,[r]δj,[r]dij∂i∂j[pr(xr, t)
= δi,[r]δj,[r]dij∂i∂jpr(xr, t)
dxlp(xlxr, t)]
(18)
An explicit calculation of the conditional averages appearing in Eq.(17) yields (cid:104)xlxr(cid:105) = C[l,r]C−1
substitute contributions (I), (II) and (III) in Eq. (15) under steady state conditions to obtain Eq. (3).
[r,r]xr [34]. We can
DERIVATION OF EQ. (5)
(cid:90)
Here we derive the expression in Eq. (5) for the cycling frequencies. To this end, we first show that the right hand
side of this equation is invariant under orientation preserving linear transformations restricted to the 2-dimensional
reduced subspace. Let us consider such a transformation: x(cid:48)
[r,r] the reduced
covariance matrix in the transformed coordinates.
r = Bfr, and denote by C(cid:48)
r = Bxr, f(cid:48)
(cid:90)
(cid:90)
(cid:90)
(16)
(17)
(19)
(20)
Using this result together with the transformation properties of the vector product, we obtain
BC[r,r]BT = C(cid:48)
[r,r] =⇒ det B =
(cid:104)x(cid:48)
r × f(cid:48)
(cid:112)det C[r,r]
r(x(cid:48))(cid:105)
=
1
det B
=
(cid:11)(cid:113)
(cid:10)τ(cid:48)
det C(cid:48)
ij
.
[r,r]
(cid:104)τij(cid:105)(cid:112)det C[r,r]
=
(cid:104)xr × fr(x)(cid:105)
(cid:112)det C[r,r]
(cid:90)
(cid:115)
det C(cid:48)
[r,r]
det C[r,r]
(cid:90)
1
γ
The coordinate invariance of this term allows us to specifically consider the convenient coordinates in which C[r,r] = I:
(cid:104)τij(cid:105) =
1
γ
1
γ
(cid:104)xr × fr(x)(cid:105) =
1
γ
dxr (cid:104)xr × fr(x)xr(cid:105) pr(xr) =
dxr xr × (cid:104)fr(x)xr(cid:105) pr(xr)
(21)
We can further expand this expression by using Ωr = Aeff + D[r,r]C−1
from Eq. (4), since we require jr(xr) = Ωrxrpr(xr).)
[r,r]. (The expression for Ωr follows immediately
7
(cid:104)fr(x)xr(cid:105) = Aeffxr = Ωrxr − D[r,r]C−1
[r,r]xr.
1
γ
(cid:90)
(cid:90)
Combining this result with Eq. (21), we arrive at
(cid:104)τij(cid:105) =
1
γ
dxr xr × (Ωrxr)pr(xr) −
dxr xr × (D[r,r]C−1
[r,r]xr)pr(xr).
Using the explicit form of Ωr (see Eq. (31)), we evaluate the first term in this expression,
dxr xr × (Ωrxr)pr(xr) =
dxr ωij(x2
i + x2
j )pr(xr) = ωij(cii + cjj) = 2ωij.
In addition, we confirm by direct calculation, that, as expected, the second term in Eq. (23) vanishes:
(cid:90)
(cid:90)
(cid:90)
(cid:90)
(cid:90)
= cij(cid:124)(cid:123)(cid:122)(cid:125)
=
0
(cid:18) dii dij
(cid:19)(cid:18)xi
(cid:19)
dij djj
dxr (−xj, xi)
dxr [−diixixj − dijx2
j + dijx2
(djj − dii) + dij (cii − cjj)
xj
= 0
(cid:124)
(cid:123)(cid:122)
0
(cid:125)
pr(xr) =
i + djjxixj]pr(xr) =
−
dxr xr × (D[r,r]xr)pr(xr) =
Altogether, this gives us the desired result:
(cid:104)τij(cid:105)(cid:112)det C[r,r]
1
2γ
= ωij
DERIVATION OF EQ. (13)
(22)
(23)
(24)
(25)
(26)
(27)
(28)
(cid:21)
p(xr)
(29)
Here we show that Πtot ≥ Πrr.
Πtot − Πr
jT (x)D−1j(x)
(cid:90)
=
dx
(cid:90)
−
dxr
kB
p(x)
(cid:90)
j∈[l]
(cid:90)
(cid:88)
(cid:88)
(cid:88)
j∈[l]
=
=
γ
kB
γ
kB
=
γ
kB
dx
v2
j (x)
(T + αj)
p(x) +
dx
v2
j (x)
(T + αj)
(cid:104)v2
j (x)(cid:105)
(T + αj)
+
p(x) +
(cid:90)
(cid:88)
p(x)
dx
i∈[r]
γ
kB
v2
i (x)
(T + αi)
[r,r]jr(xr)
(cid:20)(cid:18)(cid:90)
r (xr)D−1
jT
(cid:88)
p(xr)
(cid:90)
(cid:88)
i (x)xr(cid:105) − (cid:104)vi(x)xr(cid:105)2(cid:1)
(cid:0)(cid:104)v2
(cid:124)
(cid:125)
(cid:20)(cid:18)(cid:90)
(cid:123)(cid:122)
v2
i (x)
i∈[r]
dxr
dxl
dxr
(T + αi)
(cid:19)
(cid:90)
−
(cid:21)
p(xr)
(cid:104)vi(x)xr(cid:105)2
(T + αi)
dxr
p(xlxr)p(xr)
− (cid:104)vi(x)xr(cid:105)2
(T + αi)
(cid:19)
≥ 0
p(xr)
where in the second line we use that D is diagonal, v(x) = j(x)/p(x), and jr(xr) = p(xr)(cid:82) dxl vr(x)p(xlxr) =
(T + αi)
i∈[r]
j∈[l]
≥0
p(xr)(cid:104)vr(x)xr(cid:105), which follows from the derivation of Eq. (3).
Here we derive the expression for the partial entropy production rate in terms of the cycling frequencies (see Eq.(14)).
It is convenient to substitute the current field j = Ωxp(x) in Eq. (11), which gives
ij(D−1)jlΩlmxm
(30)
DERIVATION OF EQ. (14)
(cid:90)
dx(Ωx)T D−1(Ωx)p(x) = kB
dx xiΩT
ij(D−1)jlΩlmcmi = kB Tr (ΩT D−1ΩC).
(cid:90)
Π = kB
= kBΩT
Since the entropy production is invariant under coordinate transformations, we can use a more suitable coordinate
system. In particular, we choose a set of coordinates such that C = 1. In this set of coordinates, the entries of the
matrix Ωij correspond to the cycling frequencies in the coordinates space of the ith and jth coordinates [13]. Thus,
in the 2D case Ωr is given by
Furthermore, in this coordinate system C[r,r] and Ωr commute, yielding
r = kBω2 Tr (C[r,r]D−1
Π2D
[r,r])
Note, this expression is invariant under coordinate transformations.
(cid:19)
(cid:18) 0 ω−ω 0
Ωr =
8
(31)
(32)
|
1503.05474 | 3 | 1503 | 2015-10-20T15:53:41 | Voltage dependence of rate functions for Na+ channel inactivation within a membrane | [
"physics.bio-ph",
"q-bio.NC"
] | The inactivation of a Na+ channel occurs when the activation of the charged S4 segment of domain DIV is followed by the binding of an intracellular hydrophobic motif which blocks conduction through the ion pore. The voltage dependence of Na+ channel inactivation is, in general, dependent on the rate functions of the S4 sensors of each of the domains DI to DIV. If the activation of a single voltage sensor that regulates the Na+ channel conductance is coupled to a two-stage inactivation process, the rate functions for inactivation and recovery from inactivation, as well as the time dependence of the Na+ current in terms of the variables m(t) and h(t), may be derived from a solution to the master equation for interdependent activation and inactivation. The rate functions have a voltage dependence that is consistent with the Hodgkin-Huxley empirically determined expressions, and exhibit saturation for both depolarized and hyperpolarized clamp potentials. | physics.bio-ph | physics |
Voltage dependence of rate functions for Na+ channel inactivation
within a membrane
S. R. Vaccaro
Department of Physics, University of Adelaide, Adelaide, South Australia,
5005, Australia
[email protected]
Abstract
The inactivation of a Na+ channel occurs when the activation of
the charged S4 segment of domain DIV is followed by the binding of
an intracellular hydrophobic motif which blocks conduction through
the ion pore. The voltage dependence of Na+ channel inactivation
is, in general, dependent on the rate functions of the S4 sensors of
each of the domains DI to DIV. If the activation of a single volt-
age sensor that regulates the Na+ channel conductance is coupled
to a two-stage inactivation process, the rate functions for inactiva-
tion and recovery from inactivation, as well as the time dependence
of the Na+ current in terms of the variables m(t) and h(t), may
be derived from a solution to the master equation for interdepen-
dent activation and inactivation. The rate functions have a voltage
dependence that is consistent with the Hodgkin-Huxley empirically
determined expressions, and exhibit saturation for both depolarized
and hyperpolarized clamp potentials.
1
INTRODUCTION
The opening and subsequent inactivation of Na+ channels and the activation
of K+ channels generate the action potential in nerve and muscle membranes [1].
The Na+ channel transient current during a depolarizing voltage clamp may be
described by the expression m3h where the activation variable m and inactiva-
tion variable h satisfy first order rate equations with rate functions dependent on
the potential difference across the membrane. Support for the assumption that
activation and inactivation are separate processes was provided by the removal
of Na+ inactivation from the squid axon membrane by the internal perfusion
of pronase without affecting activation kinetics [2]. However, there is a delay
in the onset of Na+ channel inactivation that is dependent on the time-course
of channel activation, and Na+ channel inactivation partially immobilizes the
gating charge associated with activation, and it was assumed that the volt-
age dependence of inactivation was derived from the Na+ activation process
[3]. There is also a delay in the recovery from inactivation that is dependent
on the time-course of deactivation, and the rate of recovery from inactivation
saturates for large hyperpolarizing potentials [4], and therefore, activation and
inactivation are interdependent or coupled processes.
The Na+ channel protein is comprised of four domains DI to DIV, each
containing six alpha-helical segments S1 to S6, and in each domain the voltage
sensor, the S4 segment, has positively charged residues located at every third
position. The re-entrant loop between S5 an S6 forms the ion-selective filter
at the extracellular end of the pore, whereas the intracellular end of the pore
is formed by the S6 segments. The inactivation gate is an IMF motif that is
positioned on an intracellular loop between DIII and DIV, and interacts with
and blocks the flow of ions through the inner mouth of the pore [5]. Based on
voltage clamp fluorometry, in response to membrane depolarization, the trans-
verse motion of the charged S4 segments of the Na+ channel domains DI to
DIII is associated with activation, whereas the slower movement of DIV S4 is
correlated with inactivation [6, 7]. This may occur for small depolarizations
when the ion channel is usually closed (closed-state inactivation) or for larger
depolarizations when the S4 segments of the domains D1 to D3 are activated
(open-state inactivation).
In a naturally occurring paramyotonia congenita mutation of the outermost
arginine residue of DIV S4 in the human muscle Na+ channel, the inactivation
rate is decreased with little voltage dependence for moderate depolarizations
[8, 9], and therefore, the voltage dependence of inactivation is dependent on
charged residues in the S4 segment of the DIV domain. The voltage dependence
of the open to inactivated transition was also demonstrated by comparison of
gating current measurement in wild-type and ApA toxin modified cardiac Na+
channels [10]. By measurement of the OFF gating charge during repolarization
in an inactivation modified mutant of the human heart Na+ channel, it was
estimated that the DIV S4 sensor contributes approximately 30% to the OFF
charge, approximately 20% may be attributed to the DIII S4 sensor when the
inactivation gate is intact, and the rate-limiting step is the motion of the DIV
S4 sensor and not the unbinding of the inactivation gate [11].
2
In order to account for the effect of double-cysteine mutants of S4 gating
charges on the ionic current of the bacterial Na+ channel NaChBac, it has been
proposed that at least two transitions are required during the activation of each
voltage sensor [12]. This conclusion is consistent with an earlier result that cross-
linking a DIV S4 segment from the extracellular surface inhibits inactivation
during membrane depolarization whereas cross-linking the same segment from
the inside inhibits activation of the Na+ channel, and therefore, the DIV S4
sensor translocates across the membrane in two stages [13]. A Na+ channel
model that assumes that the motion of the DIV S4 sensor includes a first stage
that is necessary for opening of the channel, and a second stage that is required
for inactivation, provides a good description of gating and ionic currents [14].
The measurement of gating currents for charge neutralized segments in each
domain of the Na+ channel gives additional support to the conclusion that
the two stage activation of the DIV S4 sensor is correlated with ion channel
inactivation [15].
In this paper, expressions for the voltage dependence of the rate of inactiva-
tion and recovery from inactivation are derived by assuming that Na+ channel
inactivation is a two stage process, where the activation of DIV S4 is correlated
with the binding of the inactivation motif to the ion pore. However, Na+ chan-
nel inactivation is, in general, dependent on the rate functions of the S4 sensors
of each of the domains DI to DIV, and from a solution of the master equation
for the activation of Na+ channel conductance by a single voltage sensor that
is coupled to a two-stage inactivation process, it is shown that the voltage de-
pendence of the rate functions for inactivation and recovery from inactivation
have a similar form to empirical expressions for Na+ channels [1, 4], and in par-
ticular, the exponential variation exhibits saturation for both depolarized and
hyperpolarized clamp potentials.
INDEPENDENT ACTIVATION AND INACTIVATION OF A
Na CHANNEL
Inactivation of a Na+ channel may be described as the transverse motion of
the charged S4 segment of the domain DIV, with rate functions αi and βi, fol-
lowed by the binding of an intracellular hydrophobic motif which blocks conduc-
tion through the ion pore, with rate functions γi and δi (see Fig. 1). Assuming
that the transition of the DIV S4 segment across two potential barriers occurs
within an energy landscape, it may be shown from a solution of the Smolu-
chowski equation [16, 17, 18] that the occupation probabilities of the permissive
states h1, h2 and the inactivated state h3 are determined by
dh1
dt
dh2
dt
dh3
dt
= −αih1(t) + βih2(t),
= αih1(t) + δih3(t) − (βi + γi)h2(t),
= γih2(t) − δih3(t),
(1)
(2)
(3)
where each rate is an exponential function of the membrane potential. If the
3
Na+ channel is depolarized to a clamp potential V from a large hyperpolarized
holding potential (h1(0) = 1, and h2(0) = h3(0) = 0), and it is assumed that
the first forward and backward transitions are rate-limiting [19, 20] ( βi ≫ δi
and γi ≫ αi), the solution of Eqs. (1) to (3) is [21]
h3I (t) =
αiγi
ω1ω2
+
αiγi
ω1(ω1 − ω2)
exp(−ω1t) −
αiγi
ω2(ω1 − ω2)
exp(−ω2t).
(4)
where ω1 ≈ (γiαi + δi(αi + βi))/(γi + βi) and ω2 ≈ γi + βi ≫ ω1.
However, if the DIV S4 sensor is initially in the inactivated state (h1(0) =
h2(0) = 0, and h3(0) = 1), and the membrane is hyperpolarized to a potential
V , it may be shown that
h3D(t) =
αiγi + δi(αi + βi) exp(−ω1t)
αiγi + δi(αi + βi)
.
Eqs. (4) and (5) are solutions of the rate equation [1]
dh3
dt
= βh,2 − (αh,2 + βh,2)h3,
where
αh,2(V ) ≈
βh,2(V ) ≈
,
δi(αi + βi)
γi + βi
αiγi
.
γi + βi
Therefore, the probability of the permissive states h = h1 + h2 satisfies
dh
dt
= αh,2 − (αh,2 + βh,2)h,
(5)
(6)
(7)
(8)
(9)
and Eqs.
functions αh and βh for the squid axon Na+ channel [1] (see Fig. 2).
(7) and (8) provide a good fit to the empirical inactivation rate
Based on the measurement of a rising phase of the gating current in a squid
axon membrane and the chemical structure of a Na channel, at least two tran-
sitions are required for the activation of each voltage sensor [12, 22]. Therefore,
for each voltage sensor from domains DI to DIII, assuming no cooperativity
between sensors, the occupation probabilities of the closed states m1, m2 and
the open state m (see Fig. 4) are determined by
dm1
dt
dm2
dt
dm
dt
= −αam1(t) + βam2(t),
= αam1(t) + δam(t) − (βa + γa)m2(t),
= γam2(t) − δam(t).
(10)
(11)
(12)
where the transition rates αa, βa, γa and δa are exponential functions of the
membrane voltage V . If we assume that βa ≫ δa and γa ≫ αa [19, 20], from the
4
solution of Eqs. (10) to (12) during activation (m1(0) = 1), and deactivation
(m(0) = 1), it may be shown that the open state m may be approximated by
mA(t) ≈
mD(t) ≈
αaγa
[1 − exp(−ω1t)],
αaγa + δa(αa + βa)
αaγa + δa(αa + βa) exp(−ω1t)
.
αγa + δa(αa + βa)
(13)
(14)
where the low frequency ω1 ≈ (γaαa + δa(αa + βa))/(γa + βa). Eqs. (13) and
(14) satisfy the rate equation [1]
dm
dt
= αm,2 − (αm,2 + βm,2)m,
and the rate functions
αm,2(V ) ≈
βm,2(V ) ≈
αa
,
1 + βa/γa
δa(αa + βa)
γa + βa
,
(15)
(16)
(17)
provide a good fit to the empirical functions αm and βm for the squid axon Na
channel [1] (see Fig. 4). However, the cooperativity between the S4 sensors
in domains DI to DIII also contributes to the voltage dependence of the Na+
conductance rate functions, and more recent models have adopted exponential
functions for both αm and βm [4].
COUPLED MODELS OF ACTIVATION AND INACTIVATION
OF A Na CHANNEL
The time-dependence of the Na+ current in the squid axon may be expressed
as m3h where the activation variable m(t) and inactivation variable h(t) satisfy
the rate equations [1]
dm
dt
dh
dt
= αm − (αm + βm)m,
= αh − (αh + βh)h.
(18)
(19)
The Hodgkin-Huxley (HH) description of the Na current is equivalent to an
8-state master equation where three independent voltage sensors may activate,
and inactivation may occur from the open state or from each of the closed states
[23]. In this section, we assume that the activation of a single voltage sensor
regulating the Na channel conductance is coupled to a two-stage inactivation
process (see Fig. 5), and therefore, the kinetics may be described by a master
equation where the occupation probabilities of the closed states C1 and A1, the
open states O and A2, and the inactivated (or blocked) states B1 and B2 are
determined by
dC1
dt
= −(α1 + αO)C1(t) + βOO(t) + β1A1(t)
(20)
5
dO
dt
dA1
dt
dA2
dt
dB1
dt
dB2
dt
= αOC1(t) − (βO + α2)O(t) + β2A2(t)
(21)
= α1C1(t) − (αA + β1 + γ1)A1(t) + δ1B1(t) + βAA2(t)
(22)
= α2O(t) − (βA + β2 + γ2)A2(t) + δ2B2(t) + αAA1(t)
(23)
= γ1A1(t) − (αB + δ1)B1(t) + βBB2(t)
= γ2A2(t) + αBB1(t) − (βB + δ2)B2(t),
(24)
(25)
and the transition rates are functions of the membrane voltage V .
Assuming that the first forward and backward transitions for inactivation
are rate limiting, βj ≫ δj and γj ≫ αj , for j = 1, 2, A1 and A2 satisfy
A1 ≈
A2 ≈
,
α1C1 + δ1B1
β1 + γ1
α2O + δ2B2
β2 + γ2
,
(26)
(27)
and therefore, Eqs. (20) to (25) may be reduced to a four state master equation
(see Fig. 6)
dC1
dt
dO
dt
dB1
dt
dB2
dt
= −(ρ1 + αO)C1(t) + βOO(t) + σ1B1(t)
= αOC1(t) − (βO + ρ2)O(t) + σ2B2(t)
= ρ1C1(t) − (αB + σ1)B1(t) + βBB2(t)
= ρ2O(t) + αBB1(t) − (βB + σ2)B2(t),
where the derived inactivation rate functions
ρ1 ≈
σ1 ≈
α1γ1
, ρ2 ≈
β1 + γ1
δ1(β1 + α1)
α2γ2
β2 + γ2
,
β1 + γ1
, σ2 ≈
δ2(β2 + α2)
β2 + γ2
.
(28)
(29)
(30)
(31)
(32)
(33)
If Na+ activation and inactivation are independent processes (αO = αB,
βO = βB, ρ1 = ρ2 = ρ and σ1 = σ2 = σ), and the Na+ channel is depolarized
to a clamp potential V from a large hyperpolarized holding potential, the open
state O(t) = m(t)h(t) where the activation and inactivation variables
αO
(1 − exp[−(αO + βO)t]) ,
m(t) =
h(t) =
αO + βO
σ + ρ exp(−(ρ + σ)t)
,
(34)
(35)
σ + ρ
6
and satisfy the rate equations (18) and (19) when αO = αm, βO = βm, ρ = βh
and σ = αh (see Fig. 7). Similarly, if the Na+ channel is hyperpolarized
to a clamp potential V from a large depolarized holding potential, C1(t) =
(1 − m(t))h(t) (see Fig. 8) where
m(t) =
h(t) =
αO + βO exp(−(αO + βO)t)
αO + βO
(1 − exp[−(ρ + σ)t]).
σ
ρ + σ
,
(36)
(37)
In the HH model, the inactivation rate functions are assumed to be indepen-
dent of the Na+ conductance activation functions but, by contrast, the rate of
recovery for inactivation in hippocampal neurons is also dependent on the Na+
channel deactivation functions [4].
More generally, the Na+ current during deactivation of the channel is very
small [3], and therefore, the deinactivation rate σ2 ≈ 0. For a depolarizing
clamp potential V from a large hyperpolarized holding potential, the solution
of Eqs. (28) to (31) when ρ1 = ρ2 = ρ is
C1(t) = C1s + Σ3
O(t) = Os − Σ3
B1(t) = B1s + Σ3
B2(t) = B2s + Σ3
j=1kj+1(ωj − βO − ρ2) exp(−ωjt)
j=1kj+1αO exp(−ωjt)
j=1kj+1X1(ωj) exp(−ωjt)
j=1kj+1X2(ωj) exp(−ωjt),
(38)
(39)
(40)
(41)
where C1s = k1σ1βB(βO + ρ), Os = k1σ1βBαO, B1s = k1ρβB(αO + βO + ρ),
B2s = k1ραB(αO + βO + ρ) + ραOσ1, k−1
1 = [(αO + βO + ρ)(σ1βB + ρ(αB +
βB)) + ρσ1αO],
k2 =
k3 =
k4 =
X1(ω) =
X2(ω) =
1 − k1σ1βBω2 + k4(ω2 − ω3)
ω1 − ω2
1 − k1σ1βBω1 + k4(ω1 − ω3))
ω2 − ω1
−B1s(ω2 − ω1) + r2X1(ω1) − r1X1(ω2)
(ω2 − ω1)X1(ω3) + (ω3 − ω2)X1(ω1) − (ω3 − ω1)X1(ω2)
−ραOβB − ρ(ω − βB)(ω − βO − ρ)
ω2 − ω(αB + βB + σ1) + σ1βB
,
−ραB(αO + βO + ρ) − ραOσ1 + ρω(αO + αB)
ω2 − ω(αB + βB + σ1) + σ1βB
.
(42)
(43)
, (44)
(45)
(46)
For depolarization clamp potentials, ω2 ≈ αO + βO + ρ, ω3 ≈ αB + βB + σ1, and
ω1 ≈ αh + βh,
(47)
where the inactivation rate
βh =
ρ
1 + σ1/(αB + βB)
+
ρσ1αO
(αO + βO + ρ)(αB + βB + σ1)
,
(48)
7
and the rate of recovery from inactivation
αh =
βBσ1
αB + βB + σ1
.
(49)
The second term in Eq.(48) only makes a contribution to βh for small clamp
potentials. If α1 = α2 and δ1 are independent of V , ω1 saturates for both large
positive and negative clamp potentials (see Fig. 9). From Eq. (39), as k4 ≈ 0
for a depolarizing potential, we may write
O(t) ≈
αO
αO + βO (cid:18) αh(1 − exp[−(αO + βO + ρ)t])
αh + βh
+
βh exp[−(αh + βh)t](1 − exp[−(αO + βO)t])
αh + βh
(50)
(51)
(cid:19) ,
and therefore, O(t) ≈ m(t)h(t) (see Fig. 10) where m(t) is defined in Eq. (34)
and
αh + βh exp(−(αh + βh)t)
h(t) =
αh + βh
.
(52)
That is, the HH description of the Na+ current in terms of the variables m(t)
and h(t) is an approximation to an expression that may be derived from a
solution to a coupled model of Na+ activation and two stage inactivation for
which the the deinactivation rate σ2 ≈ 0.
For a moderate hyperpolarizing clamp potential from a depolarized holding
potential, the inactivation rates ρ1, ρ2 ≈ 0, and the solution of Eqs. (28) to (31)
is (see Fig. 11)
Y1(ω2)
ω1 − ω2
exp(−ω2t) +
exp(−ω3t)
(53)
C1(t) =
O(t) =
βO
+
Y1(ω1)
ω1 − ω2
αO + βO
αOω1ω2 + ω3σ2(βO − αB − σ1)
exp(−ω1t) −
ω3(ω1 − ω3)(ω2 − ω3)
αO
+
Y2(ω1)
ω1 − ω2
αO + βO
αOω1ω2 + ω3σ2(βO − αB − σ1)
exp(−ω1t) −
B1(t) = −
ω3(ω1 − ω3)(ω2 − ω3)
βB
exp(−ω1t) +
βB
ω1 − ω2
ω1 − ω2
Y2(ω2)
ω1 − ω2
exp(−ω2t) −
exp(−ω3t),
exp(−ω2t),
(54)
(55)
B2(t) =
ω1 − αB − σ1
ω1 − ω2
exp(−ω1t) −
ω2 − αB − σ1
ω1 − ω2
exp(−ω2t),
(56)
where ω3 = αO + βO, and ω1, ω2 are solutions of the characteristic equation
ω2
− ω(αB + βB + σ1 + σ2) + σ1βB + σ2(αB + σ1) = 0,
(57)
Y1(ω) =
Y2(ω) =
−βOω1ω2 + ω(σ1βB + σ2βO)
ω(ω − αO − βO)
,
ω1ω2(σ2 − αO) − ωσ2(βB + σ2 − αO)
ω(ω − αO − βO)
.
(58)
(59)
8
If ρ1 = ρ2, from the application of microscopic reversibility to the four state
system (see Fig. 6), βB = βOαBσ2/αOσ1 ≪ βO for σ1 ≫ σ2 ≈ 0. From
Eq. (57), assuming that αB ≪ βB ≪ σ1 for a small hyperpolarizing potential,
the lowest frequency ω1 ≈ βB, whereas for βB ≫ σ1, ω1 ≈ σ1 (see Fig. 9).
The conclusion that, for small hyperpolarizing potentials, the recovery rate for
inactivation αh ≈ βB ∝ βm is supported by the HH data where αh(V ) and
βm(V ) have a similar voltage dependence and βm(V ) ≈ 57αh(V ) [1]. However,
if αB is an exponential function of V such that αB + βB ≫ σ1 [4] (see Fig. 12),
from Eq. (57), ω2 ≈ αB + βB + σ1, and
ω1 ≈
σ1βB
αB + βB + σ1
.
(60)
Therefore, the voltage dependence of the rate of recovery from inactivation
is determined by the deinactivation rate σ1 and the Na+ conductance de-
activation functions [4]. For small hyperpolarizing potentials (βB ≪ αB),
ω1 ≈ σ1βB/(αB + σ1), and may be approximated by an exponential function of
V [1] but saturates at a more negative potential when βB ≫ σ1 ≫ αB (see Fig.
12).
From Eq. (53), we may write
C1(t) ≈
βO
αO + βO (cid:18)1 − exp(−ω1t)(cid:20)1 +
ω1(1 − exp[−(ω2 − ω1)t])
ω2 − ω1
(cid:21)(cid:19) .
(61)
and therefore, dC1/dt(0) = 0 and there is a delay in the recovery from inacti-
vation [4, 15]. However, for large negative potentials, ω2 ≈ βB ≫ ω1 ≈ σ1, and
Eq. (61) reduces to the HH expression
C1(t) ≈ [1 − ms]h(t),
(62)
where ms = αO/(αO + βO) and h(t) = 1 − exp(−ω1t).
CONCLUSION
Hodgkin and Huxley described the voltage dependence of the Na+ channel
inactivation rate and the rate of recovery by exponential functions which for
the inactivation rate saturates for a moderate depolarizing potential [1]. Based
on the absence of a gating current that corresponds to the time course of the
inactivated Na+ current, it was assumed that the transition rates governing
inactivation were voltage-independent, and that the macroscopic voltage de-
pendence of inactivation derived from the Na+ channel activation process [3].
However, the voltage dependence of the inactivation rate is dependent on the
charge on the S4 segment residues of the DIV domain [8, 9, 10], and gener-
ally, only has a minor contribution from the Na+ conductance activation rate
functions. The voltage dependence of the rate of recovery from inactivation
saturates for a large hyperpolarizing potential, and has been attributed to the
Na+ channel deactivation rate functions [4].
In this paper, assuming that Na+ channel inactivation is a two stage process,
where the activation of DIV S4 is correlated with the binding of the inactivation
9
motif to the ion pore, we show that during a voltage clamp of the Na+ channel,
the solution of the master equation for the inactivation process may be approx-
imated by the solution of a rate equation. The backward transition rate is, in
general, an exponential function of the membrane potential V , and the forward
rate may be expressed as an exponential function for small depolarizations but
approaches a saturated value for a larger clamp potential, reflecting the voltage
independence of the rate limiting step.
If the processes of Na activation and inactivation are independent, and the
activation of a single voltage sensor that regulates the Na channel conductance
is coupled to a two-stage inactivation process, the open state probability O(t)
during a depolarizing clamp potential, derived from an analytical solution of
a four state master equation, may be expressed as m(t)h(t) where m(t) and
h(t) satisfy rate equations for activation and inactivation, and the voltage de-
pendence of the rate of inactivation provides a good approximation to the HH
function βh [1]. However, the voltage dependence of the rate of recovery from
inactivation is dependent on the rate functions for the DIV sensor and not the
Na+ channel conductance deactivation functions, as observed experimentally in
hippocampal neurons [4].
Based on the measurement of a very small ion channel current during de-
activation [4], the Na+ channel deactivates before recovery from inactivation,
and hence the deinactivation rate σ2 to the open state is smaller than the cor-
responding rate σ1 from a deactivated state . A more general expression for the
voltage dependence of the rate of recovery from inactivation may be determined
that is approximated by an exponential function for physiological potentials [1],
but for more negative hyperpolarizing potentials, it approaches a limiting value
equal to the deinactivation rate σ1. The inactivation rate is, in general, also
dependent on the Na+ conductance activation rate functions, as well as the rate
functions for the DIV S4 sensor, but for a moderate depolarizing potential it
reduces to a two-stage expression when the DIV S4 activation rate ρ1 = ρ2.
We conclude that general expressions for the voltage dependence of rate
functions for inactivation and recovery from inactivation may be determined
from a master equation for interdependent Na+ channel activation and inac-
tivation, and are consistent with the empirical data from the squid axon [1],
hippocampal neurons [4] and Nav1.4 ion channels [15]. When Na+ conductance
activation is regulated by a single voltage sensor, the HH description of the Na+
current in terms of the activation and inactivation variables m(t) and h(t), is an
approximation to an expression derived from the solution to a coupled model of
Na+ activation and two-stage inactivation where the deinactivation rate to the
open state σ2 ≈ 0.
10
References
[1] A.L. Hodgkin and A.F. Huxley, J. Physiol. 117, 500 (1952).
[2] C. M. Armstrong, F. Bezanilla and E. Rojas, J. Gen. Physiol. 62, 375
(1973).
[3] C. M. Armstrong and F. Bezanilla, J. Gen. Physiol. 70, 549 (1977).
[4] C-C. Kuo and B.P. Bean, Neuron 12, 819 (1994).
[5] J.W. West, D.E. Patton, T. Scheuer, Y. Wang, A.L. Goldin and W.A.
Catterall, PNAS 89, 109 (1992).
[6] A. Cha, P.C. Reubens, A.C. George and F. Bezanilla, Neuron 22, 73 (1999).
[7] B. Chanda and F. Bezanilla, J. Gen. Physiol. 120, 629 (2002).
[8] M. Chahine, A.L. George Jr., M. Zhou, S. Ji, W. Sun, R.L. Barchi and R.
Horn, Neuron 12, 281 (1994).
[9] L. Q. Chen, V. Santarelli, R. Horn and R.G. Kallen, J. Gen. Physiol. 108,
549 (1996).
[10] M.F. Sheets and D.A. Hanck, J. Gen. Physiol. 106, 617 (1995).
[11] M.F. Sheets, J.W. Kyle and D.A. Hanck, J. Gen. Physiol. 115, 609 (2000).
[12] P. G. DeCaen, V. Yarov-Yarovoy, E. M. Sharp, T. Scheuer and W. A.
Catterall, Proc. Natl. Acad. Sci. USA 106, 22498 (2009).
[13] R. Horn, S. Ding and H.J. Gruber, J. Gen. Physiol. 116, 461 (2000).
[14] C. M. Armstrong, PNAS 103, 17991 (2006).
[15] D.L. Capes, M.P. Goldschen-Ohm, M. Arcisio-Miranda, F. Bezanilla and
B. Chanda, J. Gen. Physiol. 142, 101 (2013).
[16] S.R. Vaccaro, Phys. Rev. E 76, 011923 (2007).
[17] S.R. Vaccaro, J. Chem. Phys. 132, 145101 (2010).
[18] S.R. Vaccaro, J. Chem. Phys. 135, 095102 (2011).
[19] J.L. Lacroix, F.V. Campos, L. Frezza, and F. Bezanilla, Neuron 79, 8651
(2013).
[20] J.L. Lacroix and F. Bezanilla, PNAS 108, 6444 (2011).
[21] S.R. Vaccaro, Phys. Rev. E 90, 052713 (2014).
[22] R. D. Keynes, Proc. R. Soc. Lond. B 240, 425 (1990).
[23] C. M. Armstrong, Physiol. Revs. E 61, 644 (1981).
11
Αi
Γi
h1
F
h2
F
h3
Βi
∆i
Figure 1:
Inactivation model of a Na+ ion channel, where the occupation
probabilities of the permissive states h1(t), h2(t), and the inactivated state
h3(t) satisfy a master equation, and the rate functions αi, βi, γi and δi, are
voltage-dependent rate functions between states.
1
-
s
m
e
t
a
r
1
0.8
0.6
0.4
0.2
0
- 75
- 50
- 25
0
25
50
V mV
Figure 2: The derived rate functions αh,2(V) and βh,2(V) (solid line) in Eqs.
(7) and (8) provide a good approximation to the HH rate functions (ms−1)
αh = 0.07 exp(−(V + 50)/20) and βh = 1/(1 + exp(−(20 + V )/10)) (dotted line)
when the inactivation rate functions are αi(V ) = 1 ≪ γi(V ) = exp(3), and
βi(V ) = exp(−2.5(V − 10)/25) ≫ δi(V ) = 0.07 exp(−(V + 50)/20).
Αa
Γa
m1
F
m2
F
m
Βa
∆a
Figure 3: Activation model of a Na+ channel, where the occupation probabili-
ties of the closed states m1, m2, and the open state m satisfy a master equation,
and αa, βa, γa and δa are voltage-dependent rate functions between states.
12
1
-
s
m
e
t
a
r
10
8
6
4
2
0
- 50
- 25
0
25
50
V mV
Figure 4: The derived rate functions αm,2(V ) and βm,2(V ) (solid line) in Eqs.
(16) and (17) provide a good approximation to the HH rate functions (ms−1)
αm = 0.1(V +25)/(1+exp[−0.1(V +25)]) and βm = 4 exp[−(V +50)/18] (dotted
line) when the activation rate functions are αa(V ) = 2.7 exp(0.27(V +50)/25) ≪
γa(V ) = 30 exp(0.25(V + 50)/25), and βa(V ) = 251 exp(−0.95(V + 50)/25) ≫
δa(V ) = 4.65 exp(−0.9(V + 50)/18).
13
C1
ΑO
ΒO
O
Β1
Α1
Β2
Α2
A1
ΑA
ΒA
A2
∆1
Γ1
∆2
Γ2
B1
ΑB
ΒB
B2
Figure 5: Six state system that describes Na+ conductance activation between
states C1 and O, A1 and A2, and B1 and B2, is coupled to a two-stage Na+
inactivation process between states C1 and B1, and O and B2.
14
C1
ΑO
ΒO
O
Σ1
Ρ1
Σ2
Ρ2
B1
ΑB
ΒB
B2
Figure 6: The six state system of Fig. 5 may be approximated by a four state
system when βj ≫ δj and γj ≫ αj, for j = 1, 2 where ρj and σj are derived
rate functions for a two-stage Na+ inactivation process, defined in Eqs. (33)
and (33).
15
V = 60 mV
V = -10 mV
V = -30 mV
t
O
1
0.8
0.6
0.4
0.2
0
0
1
2
3
4
5
6
tms
Figure 7: For a four state system where activation and inactivation are inde-
pendent, the open state probability O(t) (solid line) during a depolarizing clamp
potential, is equal to m(t)h(t) (dashed, dotted or dot-dashed line) where m(t)
and h(t) are solutions of rate equations for activation and inactivation, and αO =
αm = 0.1(V + 25)/(1 − exp[−(V + 25)/10]), βO = βm = 4 exp[−(V + 50)/18],
ρ = βh = 1/(1 + exp[−(20 + V )/10]) and σ = αh = 0.07 exp[−(V + 50)/20]
(ms−1) [1].
t
1
C
1
0.8
0.6
0.4
0.2
0
V = -150 mV
V = -100 mV
V =
-80 mV
0
5
10
15
20
tms
Figure 8: For a four state system where activation and inactivation are in-
dependent, the closed state probability C1(t) (solid line) during a hyperpo-
larizing clamp potential,
is equal to [1 − m(t)]h(t) (dashed, dotted or dot-
dashed line) where m(t) and h(t) are solutions of rate equations for activa-
tion and inactivation, and αO = αm = 0.1(V + 25)/(1 − exp[−(V + 25)/10]),
βO = βm = 4 exp[−(V + 50)/18], ρ = βh = 1/(1 + exp(−(20 + V )/10)) and
σ = αh = 0.07 exp(−(V + 50)/20) (ms−1).
16
e
t
a
r
n
o
i
t
a
v
i
t
c
a
n
i
4
3
2
1
0
-200
-150
-100
-50
0
50
V
Figure 9: Voltage dependence of the HH Na+ channel inactivation rate func-
tions βh = 1/(1 + exp(−(20 + V )/10)) and αh = 0.07 exp(−(V + 50)/20)
(dashed line) may be approximated by analytical expressions in Eqs.
(48)
and (49) (solid line) derived from a master equation for a four state sys-
tem where activation and two stage inactivation are interdependent, and by
the voltage dependence of the lowest frequency of the system determined nu-
merically (dotted line) where the rate functions are α1(V ) = α2(V ) = 1,
γ1(V ) = γ2(V ) = exp(3), β1(V ) = β2(V ) = exp(−2(V − 10)/25) δ1(V ) = 4.2,
δ2(V ) = 0, αO = 0.1(V + 25)/(1 − exp[−(V + 25)/10]) = αB/3 and βO =
4 exp[−(V + 50)/18] = 83βB (ms−1).
17
1
0.8
0.6
0.4
0.2
0
t
O
V = 60 mV
V = -10 mV
V = -30 mV
0
1
2
3
t
4
5
6
Figure 10:
For a four state system where activation and inactivation are
interdependent, during a depolarizing clamp potential, the open state proba-
bility O(t) (solid line) may be approximated by m(t)h(t) (dashed, dotted or
dot-dashed line) where m(t) and h(t) are solutions of rate equations for acti-
vation and inactivation, and α1(V ) = α2(V ) = 1, γ1(V ) = γ2(V ) = exp(3),
β1(V ) = β2(V ) = exp(−2(V − 10)/25) δ1(V ) = 4.2, δ2(V ) = 0, αO = 0.1(V +
25)/(1 − exp[−(V + 25)/10]) = αB/3 and βO = 4 exp[−(V + 50)/18] = 83βB
(ms−1).
t
1
C
1
0.8
0.6
0.4
0.2
0
V = -150 mV
V = -100 mV
V =
-80 mV
0
5
10
15
20
tms
Figure 11: For a hyperpolarizing clamp potential of a four state system where
activation and inactivation are interdependent, the closed state variable C1(t)
(solid line) may be approximated by [1 − ms]h(t) (dashed, dot or dot-dashed)
where ms = αO/(αO + βO) and h(t) is a solution of a rate equation for inac-
tivation, and rate functions are α1(V ) = α2(V ) = 1, γ1(V ) = γ2(V ) = exp(3),
β1(V ) = β2(V ) = exp(−2(V − 10)/25) δ1(V ) = 4.2, δ2(V ) = 0, αO = 0.1(V +
25)/(1 − exp[−(V + 25)/10]) = αB/3 and βO = 4 exp[−(V + 50)/18] = 83βB
(ms−1).
18
1
-
s
m
e
t
a
r
n
o
i
t
a
v
i
t
c
a
n
i
4
3
2
1
0
- 200
- 150
- 100
- 50
0
50
VmV
Figure 12: Voltage dependence of the HH Na+ channel inactivation rate func-
tions βh = 1/(1+exp(−(20+V )/10)) and αh = 0.07 exp(−(V +50)/20) (dashed
line) may be approximated by analytical expressions (solid line) derived from
a master equation for a four state system where activation and two stage in-
activation are interdependent, and by the voltage dependence of the lowest
frequency of the system determined numerically (dotted line) where the rate
functions are α1(V ) = α2(V ) = 1, γ1(V ) = γ2(V ) = exp(3), β1(V ) = β2(V ) =
exp(−2.5(V − 10)/25) δ1(V ) = 4.2, δ2(V ) = 0, αO = αB = 5.4 exp(−0.3V /25)
and βO = 0.5 exp[−V /18] = 98βB (ms−1).
19
|
1703.06276 | 1 | 1703 | 2017-03-18T09:07:21 | Droplet Ripening in Concentration Gradients | [
"physics.bio-ph"
] | Living cells use phase separation and concentration gradients to organize chemical compartments in space. Here, we present a theoretical study of droplet dynamics in gradient systems. We derive the corresponding growth law of droplets and find that droplets exhibit a drift velocity and position dependent growth. As a consequence, the dissolution boundary moves through the system, thereby segregating droplets to one end. We show that for steep enough gradients, the ripening leads to a transient arrest of droplet growth that is induced by an narrowing of the droplet size distribution. | physics.bio-ph | physics |
Droplet Ripening in Concentration Gradients
Christoph A. Weber1,2, Chiu Fan Lee3 and Frank Julicher1,2
1 Max Planck Institute for the Physics of Complex Systems, Nothnitzer Str. 38,
01187 Dresden, Germany
2 Center for Advancing Electronics Dresden cfAED, Dresden, Germany
3 Department of Bioengineering, Imperial College London, South Kensington
Campus, London SW7 2AZ, U.K.
E-mail: [email protected], [email protected]
Abstract. Living cells use phase separation and concentration gradients to organize
chemical compartments in space. Here, we present a theoretical study of droplet
dynamics in gradient systems. We derive the corresponding growth law of droplets
and find that droplets exhibit a drift velocity and position dependent growth. As a
consequence, the dissolution boundary moves through the system, thereby segregating
droplets to one end. We show that for steep enough gradients, the ripening leads to a
transient arrest of droplet growth that is induced by an narrowing of the droplet size
distribution.
1. Introduction: Droplet ripening in concentration gradients in biology
Living cells have to organize many molecules in space and time in order to build
compartments which can perform certain biological functions. The formation of these
compartments is often regulated by spatially heterogenous distributions of molecular
species. An example is the polarized distribution of polarity proteins in the course
of asymmetric cell division [1, 2, 3]. During asymmetric cell division, molecules of
the cell cytoplasm are distributed unequally between both daughter cells [4, 5]. This
can be studied in the first division of the fertilized egg of the roundworm C.elegans.
RNA-protein aggregates called P-granules are segregated to the posterior side of the
cell and are located in the posterior daughter cell after division. P-granules are liquid
like droplets that form by phase separation from the cell cytoplasm [6, 1, 2, 3]. The
segregation and ripening of P-granule droplets toward the posterior is driven by a
concentration gradient of the protein Mex-5 that regulates droplet dynamics [6, 7, 8].
The ripening of drops guided by a concentration gradient of molecules that regulate
phase separation fundamentally differs from classical Ostwald-ripening. In the case of
Ostwald ripening, droplets are uniformly distributed throughout the system and the
droplet size distribution broadens with time [9, 10, 11, 12]. If a concentration gradient
Droplet Ripening in Concentration Gradients
2
Figure 1.
(a) Schematic representation of the regulation of droplet (blue dot)
formation by a regulator R which can bind to droplet material D to form the product
R:D. (b) Illustration of droplet ripening in a gradient of regulator volume fraction φR
(orange). At each time point a boundary (red dashed) divides the system into domains
of growth and shrinkage. This boundary moves to the right (red arrow) leaving a region
of dissolving drops behind. Droplets drift (black arrows) with velocity Vd.
of a regulator component is maintained, for example by sources and sinks [8], or via
position-dependent reaction kinetics [13, 14], there is a broken symmetry generating a
bias of droplet positions. Recently, droplet segregation in a concentration gradient has
been discussed using a simplified model [7]. However, the dynamics of droplets ripening
in a gradient of regulating molecules has not been explored (figure 1(a)).
In this paper we present a theoretical study of droplet ripening in a concentration
gradient of a regulator that affects phase separation. Considering a simplified theory we
extract generic physical features of droplet growth in the presence of concentration
gradients. The generic features we study here are the spatially dependent,
local
equilibrium concentration and a spatially dependent actual concentration outside the
droplets. If the distance between droplets is large, these features can be used to derive
the generic laws of droplet ripening in concentrations gradients and thereby extend
the classical theory for homogeneous systems [9, 10]. Our central finding is that a
regulator gradient leads to a drift velocity and a position dependent growth of drops
(figure 1(b)). As a consequence, a dissolution boundary moves through the system,
leaving droplets only in a region close to one boundary of the system. Using numerical
calculations supported by analytic estimates, we study the growth dynamics of droplets
in a gradient. We discover that, surprisingly, ripening is not always faster in the case
of steeper regulator gradients. Instead, a transient arrest of ripening is observed that
results from a narrowing of the droplet size distribution. Our work shows that a regulator
gradient induces a novel and rich ripening dynamics in droplet systems.
Droplet Ripening in Concentration Gradients
3
2. Local regulation of phase separation
(cid:21)
(cid:20) φT
D
νD
We use a simplified model to discuss two component phase separation that is influenced
by a regulator. We consider a system consisting of a solvent S, droplet material D and
a regulator R that can create together with the droplet material a bound state R:D. In
this model the regulator does not take part in demixing but influences phase separation
of D and S. We describe demixing by a simplified Flory-Huggins type of free-energy
density
f = kBT
ln φT
D +
φS
νS
ln φS
+ E ,
(1)
where kB is the Boltzmann constant, T is temperature and φT
D and φS denote the
total volume fraction of droplet material and solvent, respectively, with φT
D + φS = 1.
In equation (1) we neglect for simplicity the mixing entropy of the component φR:D.
We only consider interactions between droplet material D and solvent S and the
corresponding interaction energies are described by E. These simplifications do not
affect the qualitative feature of position dependent phase separation that we highlight
in this work but are useful simplification for the discussion of the relevant physics. The
molecular volumes νi connect volume fractions with concentrations ci by φi = νici. The
regulator influences phase separation by binding to droplet material,
D + R (cid:42)(cid:41) R : D .
(2)
Here we consider the case where the bound state R:D does not phase separate from
the solvent. The total volume fraction of droplet material is given by the sum of
contributions of bound and free molecules, φT
D = φD + φR:D. The binding process
between regulator and droplet material can be described by mass action with the
equilibrium binding constant K0 = cR:D/(cDcR). Using the simplification νR:D = νD,
we write K0 = φR:D/(φDcR). The interaction energy is given by E = kBT χ φDφS, where
χ is the interaction parameter. Expressing φD in terms of φT
D and considering a fast
local equilibrium of the binding reaction we find
with
D, φS) = kBT χeff φT
DφS ,
E(φT
(cid:32)
1 −
χeff = χ
(cid:33)
KφR
1 + KφR
(3)
(4)
and K = K0/νR. The function χeff describes the effective interaction between the
solvent and the total droplet material which depends on the regulator.
In the case
of a vanishing regulator concentration, φR = 0, equation (1) reduces to the original
Flory-Huggins model for binary polymer blends [15] with an interaction parameter
χeff = χ.
Increasing the concentration of the regulator leads to a decrease in the
effective interaction parameter χeff. This decrease is more pronounced if the binding
constant K is larger, which amounts to more R being bound to D. Please note that
only for large values of χeff relative to the entropic terms in equation (1) demixing can
occur (figure 2(a)).
Droplet Ripening in Concentration Gradients
4
3. Spatial organization of phase separation
To describe the spatial regulation of phase separation we consider a spatially
inhomogeneous system that is locally at thermodynamic equilibrium such that at each
position the local free energy is defined. Globally the system is maintained away from
equilibrium by an imposed position dependent regulator gradient. For simplicity, we
use a linear gradient along the x direction, φR(x) = φ0 − m · x, with x ∈ [0, L], where L
denotes the size of the system. We first look at a situation without droplets but with a
possible spatial profile φD = ¯Φ(x) of droplet material. Since the spatial concentration
profile of the regulator φR(x) is imposed, the effective interaction parameter χeff(x)
becomes a function of x. As the droplet material is also distributed in space, the
concentration at each position x corresponds to a point in the phase diagram. The
linear range x ∈ [0, L] then maps onto a line that is indicated in the phase diagram
in figure 2(a). Using the phase diagram, we can determine the position xd of the
dissolution boundary, which separates the region x < xd where the fluid mixes, from the
region x > xd in which droplets can form. For x > xd, we can then determine the local
equilibrium volume fraction Φin
eq (x) outside
of a potential droplet, which depend on position. For νS (cid:29) νD, Φin
eq is approximately
constant along x. As we will see below, choosing this simple limit allows us to focus on
the concentration field outside of the droplet. The spatial distribution of the regulator
and droplet material imply a spatially dependent supersaturation defined as
eq(x) of the droplet material inside and Φout
(x) =
¯Φ(x)
eq (x) − 1 ,
Φout
(5)
which is positive for x > xd. In the absence of droplets, the concentration field ¯Φ(x)
evolves in time satisfying a diffusion equation. If droplets are nucleated, their dynamics
of growth or shrinkage is guided by the local supersaturation (x) as well as Φin
eq and
eq (x). This droplet dynamics then in turn also influences the concentration field ¯Φ(x).
Φout
4. Dynamics of a single drop in a concentration gradient
A regulator concentration gradient generates a position-dependent supersaturation
(equation (5)), which will generically influence the spatial distribution of droplet
material ¯Φ(x).
In the following we discuss the kinetics of growth of a single droplet
eq (x), and droplet material, ¯Φ(x), are position
where the equilibrium concentration, Φout
dependent, and thereby extend the classical description of droplet growth [9, 10] to the
case of concentration gradients.
Neglecting variations of the concentration inside of the droplet, we restrict ourselves
to the concentration field outside of the single droplet, φout(r, θ, ϕ). We use spherical
coordinates centred at the droplet position x0, with r denoting the radial distance from
the centre, and θ and ϕ are the azimuthal and polar angles. The volume fraction outside
but near the droplet then obeys the steady state of a diffusion equation
∇2φout(r, θ, ϕ) = 0 .
(6)
Droplet Ripening in Concentration Gradients
5
Figure 2.
(a) Phase diagram. Interaction parameter χeff as a function of the volume
fraction of droplet material φT
D. The binodal line (blue) and the critical point (triangle)
are indicated. For a regulator concentration gradient, the positions x in the system are
mapped to a line (dashed/green line in the mixed/demixed region). At the position
x = xd this line crosses the binodal. (b) Equilibrium volume fractions outside and
inside the droplet, Φout
eq, corresponding to the binodal line in (a), are shown as
functions of position x > xd for K = 500. Parameters: m = −3 · 10−3, φ0 = 4 · 10−3,
eq and Φin
νS = 10νD.
The concentration field approaches for large r (far from the drop) a linear gradient of
the form,
φout(r, θ) = α r cos θ + β ,
(7)
lim
r→∞
where the droplet material outside ¯Φ(x) is locally characterized by the concentration
β = ¯Φ(x0) and the gradient α = ∂x
¯Φ(x0) at the droplet position x0. At the surface
r = R of a spherical droplet, the boundary condition is
φout(R, θ) = (Φout
eq + R cos(θ)∂xΦout
eq )(1 + (cid:96)c/R) .
(8)
Here, (cid:96)c = 2γνD/(kbT ) is the capillary length, γ denotes the surface tension of the
droplet. Equation (8) corresponds to the Gibbs-Thomson relation [12], which describes
the increase of the local concentration at the droplet interface relative to the equilibrium
concentration due to the surface tension of the droplet. The presence of spatial
inhomogeneities on the scale of the droplet R lead to an additional term in the Gibbs-
Thomson relation of the form R cos(θ)∂xΦout
eq . The values of α and β characterizing
the far field together with the local concentration at the droplet surface, φout(R, θ),
then determine the local rates of growth or shrinkage of the drop. Deformations of
the spherical shape of the droplet can be neglected if the surface tension is large and
concentration gradients on the scale of the droplet are small. Furthermore, we focus, for
simplicity, on the case where the Onsager cross coupling coefficient between the regulator
and droplet material is negligible and we thus ignore how the spatial distribution of
droplet material affects the maintained regulator gradient.
χeffφTD012300.51(a)x=0x=LmixeddemixedxdΦouteq(x)x/L0100.51(b)0100.51Φineq(x)xDroplet Ripening in Concentration Gradients
6
(cid:80)∞n=0 (Anrn + Bnr−n−1) Pn(cos θ), where Pn(cos θ) are the Legendre polynomials. Using
The solution to the diffusion equation (6)
the form, φout(r, θ) =
is of
the boundary conditions (7) and (8), we find
(cid:32)
(cid:33)
(cid:19)
R
r
(cid:18)
(cid:17)(cid:32)
1 −
φout(r, θ) = α cos θ
R3
r2
r −
+ β
(cid:16)
+
Φout
eq + R cos(θ)∂xΦout
eq
1 +
(cid:33) R
r
.
(9)
(cid:96)c
R
The interface of a droplet at position x0 can be expressed by a function R (θ, ϕ, t; x0). The
speed of the interface is ∂tR (θ, ϕ, t; x0) = vn(θ, ϕ; x0), where vn = (cid:126)n· (cid:126)J is the local velocity
normal to the interface and (cid:126)n denotes a surface normal. Here, (cid:126)J = ((cid:126)jin−(cid:126)jout)/(Φin
D−Φout
eq )
is the local interface velocity [12], and (cid:126)jin and (cid:126)jout denote the volume fluxes at the
droplet surface inside and outside of the drop. Since the volume fraction inside the
droplet is considered as constant and independent of the droplet position, (cid:126)jin = 0 and
(cid:126)jout = −D∇φout.
eq ) the growth
the droplet radius, R = (1/4π)(cid:82) dϕdθ sin θ R , we can calculate the growth rate of the
eq)∂rφoutr=R. With the definition of
velocity normal to the interface is vn = (D/Φin
droplet radius, dR/dt = (1/4π)(cid:82) dϕdθ sin θ ∂tR , and the net drift velocity along the x-
direction, Vd = (1/4π)(cid:82) dϕdθ sin θ (cid:126)ex · (cid:126)er ∂tR . Here, (cid:126)ex · (cid:126)er = cos θ and (cid:126)er and (cid:126)ex denote
In the limit of strong phase separation (Φin
the radial unit vector in spherical coordinates and the unit vector along the x-direction
in cartesian coordinates. Thus, the droplet radius grows as
eq (cid:29) Φout
In the presence of concentration gradients there also exists a net drift velocity with
eq (x0)
1 +
.
(10)
(cid:34)
β − Φout
(cid:34)
α − ∂xΦout
dR
dt
=
D
Φin
eqR
Vd =
D
Φin
eq
(cid:32)
(cid:32)
(cid:96)c
R
(cid:33)(cid:35)
(cid:33)(cid:35)
(cid:96)c
R
(cid:35)
eq (x0)
1 +
.
(11)
Note that both the growth speed and the drift velocity are set by the molecular
diffusion constant D of droplet material.
5. Ripening of multiple drops in a regulator gradient
We can now describe the dynamics of many droplets i = 1,··· , N , with positions xi and
radius Ri. If droplets are far apart from each other, the rate of growth of droplet i reads
(cid:34)
d
dt
Ri =
D
Ri
Φout
eq (xi)
Φin
eq
(xi) −
(cid:96)c
Ri
.
The droplet drift velocity, dxi/dt = Vd(xi), is given by
dxi
dt
=
D
Φin
eq
∂x
¯Φ(x)xi − ∂xΦout
eq (x)xi
1 +
(cid:34)
(cid:32)
(cid:33)(cid:35)
.
(cid:96)c
Ri
(12)
(13)
Droplet Ripening in Concentration Gradients
7
If the distance between droplets is large relative to their size, droplets only interact via
the concentration field ¯Φ(x, t) which represents the far field. It is governed by a diffusion
equation including gain and loss terms associated with growth or shrinkage of drops:
N(cid:88)
i=1
¯Φ(x, t) = D
∂t
∂2
∂x2
¯Φ(x, t) −
4πΦin
eq
3L3
δ(xi − x)
d
dt
R3
i (t) .
(14)
For simplicity, in the above equation we consider a regulator gradient along the x axis.
Please note that equation (14) describes the effects of large scale spatial inhomogeneities
on the ripening dynamics. Since large scale variations of ¯Φ(x, t) only build up along the
x-directions, derivatives of ¯Φ along the y and z axes do not contribute.
In the absence of a regulator gradient, Φout
eq and ¯Φ are constant implying a position-
independent supersaturation level (equation (5)). In this case equation (12) gives the
classical law of droplet ripening derived by Lifschitz-Slyozov [9, 10] (also referred to as
Ostwald-ripening), and the net drift vanishes (equation (13)). In the case of Ostwald
ripening large droplets of radius larger than the critical radius, Rc = (cid:96)c/, grow at the
expense of smaller shrinking drops. This causes an increase of the average droplet size
and a broadening of the droplet size distribution with time. On large spatial scales,
droplets remain homogeneously distributed in the system.
This property fundamentally changes due to the presence of concentration gradients
leading to two possibilities of droplet material transport along the regulator gradient:
(i) Exchange of material between droplets at different positions of the concentration
gradient by diffusive transport in the dilute phase or (ii) drift of droplets along the
(i) Droplets grow or shrink with rates that vary along the
concentration gradient.
gradients of local equilibrium volume fraction Φout
eq and the droplet material volume
fraction ¯Φ(x) (Eq. (12)). For (x) = ¯Φ(x)/Φout
eq (x) − 1 > (cid:96)c/R, a droplet located at
position x grows, and shrinks in the opposite case. The critical droplet radius thus
becomes position dependent, where below or above Rc(x) = (cid:96)c/(x) droplets shrink or
grow. (ii) The drift of a droplet (equation (13)) results from an asymmetry of material
flux though the interface parallel to the regulator gradient. If ∂x
eq , the
droplet drift velocity Vd points toward regions of smaller Φout
eq (x). This is a typical
¯Φ(x) tends to flatten with time due to the
case since the gradient of droplet material ∂x
diffusion of droplet material in the dilute phase.
¯Φ(x) < ∂xΦout
To study the ripening dynamics of droplets in a concentration gradient we solved
the equations (12) to (14) numerically. To access the late time regime of ripening
we first initialize about N = 107 drops with radii taken from the Lifschitz-Slyozov
distribution [9, 10] in a system of position independent equilibrium concentration Φ0, and
fix the concetration inside Φin
eq = 1. For t ≥ L2/D, we then spatially quench the system
by imposing the spatially varying equilibrium concentration Φout
eq (x) = Φ0(1 − s x) [16],
which we refer to as "spatial quench" in the following.
In our numerical studies we
find that droplets experience a non-uniform growth depending on the position and the
stage of ripening (figure 3(a)). At the beginning, all drops grow in the region where
the concentration ¯Φ(x) exceeds the local equilibrium concentration at the drop surface,
Droplet Ripening in Concentration Gradients
8
Figure 3.
Droplet ripening in concentration gradients. (a) Mean droplet radius
(cid:104)R(cid:105)x at position x as a function of time for different x as indicated. A spatial
profile of equilibrium volume fraction of slope s = 0.5 is imposed at time t = L2/D
(quench). The grey data points correspond to classical Ostwald ripening (s = 0). (b)
Characteristic time τL required to segregate the volume of droplet material toward
x = L, and dissolution time τD required to reach 10 droplets starting from O(104), as
a function of quench slope s. The horizontal grey line indicated the value of τD for
classical Ostwald ripening (s = 0).
Φout
eq (x) (1 + (cid:96)c/R), and shrink otherwise. The dissolution boundary at x = xd obeys
¯Φ(xd) (cid:39) Φout
eq (xd) since in the late time regime (cid:96)c (cid:28) R. It moves according to
(cid:30) dΦout
eq (x)
dx
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)x=xd(t)
dxd
dt
=
d ¯Φ(xd(t))
dt
.
(15)
For dΦout
eq (x)/dx < 0, the position of the dissolution boundary xd moves to the right
until it reaches the system boundary at x = L (Supplemental video [17]). At long
times, the volume fraction at all positions approaches the minimum of the equilibrium
eq (L), and all droplets dissolve except at x = L (figure 3(a),
volume fraction, ¯Φ(x) → Φout
Supplemental video [17]).
The characteristic time τL of droplet segregation depends on the quench slope s.
It decreases for increasing s according τL ∝ s−1 (figure 3(b)). In contrast the time of
droplet dissolution, τD, defined as the time to reach 10 droplets, changes only weakly
with the quench slope s and can even increase (figure 3(b)). Interestingly, the droplet
ripening exhibits periods of transient arrest, during which droplet number and size
remain almost constant (figure 4(a)). These arrest phases govern the time of droplet
dissolution for large quench slopes since they occur for sufficiently large quench slopes
s. The duration of arrest is roughly constant as a function of s and the onset of the
arrest phase is delayed for decreasing s [18] (figure 4(b)). Intriguingly, the onset of the
arrest phase is preceded by a narrowing of the droplet size distribution. The droplet
size distribution narrows during the segregation of droplets toward x = L while the
onset of arrest occurs after droplets have mostly been spatially segregated (Supplemental
video [17]). In particular, the standard deviation of droplet radius exhibits a pronounced
hRix/LtD/L2x=Lx=3L/4x=L/410−510−410−3100102104quench(a)τD/L2s/LτDτL10010210410−410−2100(b)∝s−1Droplet Ripening in Concentration Gradients
9
Figure 4. Narrowing droplet size distribution. (a) Mean radius (cid:104)R(cid:105) averaged over
all drops in the system as a function of time for three quench slopes s. The onset
of arrest is indicated (arrows). (b) Duration and time of onset of the arrest phase
as a function of quench slope s. The vertical black line indicates the quench slope
sc = {1− [3/2− Φ0/(2 ¯Φ(t = L2/D))]−1}/L below which no arrest can occur. It can be
calculated by the condition that the critical radius Rc at x = L is reduced by at least
a factor of 3/2 during the quench such that the largest droplets in the distribution
grow more slowly than smaller ones and the distribution narrows. For our numerical
solutions sc ≈ 0.017/L, which is consistent with the emergence of the arrest along
quench slope found in our numerical calculations. (c) Standard deviation δR of the
droplet radius distribution as a function of time for three different quench slopes s.
The onset of arrest corresponds to a sudden narrowing of the distribution (arrows).
(d) Rate of droplet growth dR/dt as a function of droplet radius R before (grey) and
after (black) the spatial quench. The droplet radius distribution p(R) at the moment
of the quench is shown (blue). Narrowing of p(R) occurs if droplet size exceeds the
radius for which the growth rate is maximal (black dots).
minimum when the arrest begins (figure 4(c)). After the arrest phase droplets undergo
classical Ostwald ripening where time-dependence of (cid:104)R(cid:105) and δR is consistent with t1/3
(figures 3(a) and 4(a)). The effect of a narrowing droplet size distribution has also
been observed in open but spatially homogeneous systems with constant influx of phase
separation material [19, 20].
The narrowing of the droplet size distribution in a concentration gradient is
hRi/LtD/L2s=0s=0.5s=0.110−410−3100102104(a)arrestonsetτD/L2s/Ldurationonset10010210410−310−210−1100noarrest(b)scarrestδR/LtD/L2s=0s=0.1s=0.510−510−4100102104(c)arrestonsetdR/dtp(R)R/Rc(s=0)-10120.1110012(d)2beforeafterDroplet Ripening in Concentration Gradients
10
fundamentally different from broadening of the droplet size distributions during classical
Ostwald-ripening [9, 10]. Ostwald ripening is characterized by a supersaturation that
decreases with time, leading to an increase of the critical droplet radius Rc = (cid:96)c/(t) ∝
t1/3. The droplet size distribution p(R) has a universal shape and is nonzero only in the
interval [0, 3Rc/2] (figure 4(d), blue graph). The broadening of p(R) follows from larger
droplets growing at a larger rate dR/dt than smaller droplets. Though dR/dt has a
maximum at R = 2Rc and decreases for large R, no droplets exist larger than 3Rc/2.
This situation changes in the presence of a concentration gradient. The spatial
quench reduces the local critical radius Rc(x (cid:39) L) = (cid:96)c/(x (cid:39) L) at the rightmost
boundary x (cid:39) L as compared to the critical radius before the quench (equation (5)).
This quench also shifts the maximum of dR/dt for droplets at x (cid:39) L to smaller radii
(black line in figure 4(d)) since the radius corresponding to the maximum occurs at
R = 2Rc. As a result, many droplets now exist after the spatial quench with large radii
R > 2Rc(x (cid:39) L). These droplets grow more slowly than those at R = 2Rc which leads to
a narrowing of the size distribution p(R) at x (cid:39) L. The critical radius Rc(x (cid:39) L) remains
small because dissolution of droplets at x < L leads to a diffusive flux toward x (cid:39) L and
thus keeps the volume fraction ¯Φ(L) at increased levels. These conditions hold longer
if the spatial quench has a steeper slope. As a result the distribution narrows more
for steeper quenches. When the critical radius catches up with the mean droplet size
narrowing stops and the onset of arrest occurs. At this time droplets have almost equal
size which slows down the exchange of material between droplets via Ostwald ripening,
leading to a long phase of almost constant size and number of droplets (figure 3(a)).
During this arrest phase, the droplet distribution broadens slowly.
6. Conclusion and Outlook
Here we presented the generic behavior of droplet ripening in concentrations gradients
and extended the classical theory by Lifschitz & Slyozov to inhomogeneous systems [9,
10]. One main result is that a concentration gradient of a soluble component that
regulates liquid-liquid phase separation can reshape the supersaturation profile such
that all drops dissolve except those within a region close to one boundary of the system.
As a consequence droplets segregate toward the boundary where the supersaturation
is highest [7].
Even though the details by which a regulator affects the local
supersaturation are system-specific, the resulting ripening dynamics that takes place
in a supersaturation gradient is generic. Surprisingly, we find that the size distribution
of droplets narrows for sufficiently steep concentration gradients, leading to a transient
arrest of the droplet dynamics. Such a behavior is fundamentally different to classical
Ostwald-ripening where the droplet size distribution continuously broadens at all times.
Transient narrowing of the droplet size distribution stems from a position-dependent
shift of the maximal droplet growth rate to smaller droplet radii as compared to spatially
homogeneous systems (figure 4(d)).
Our work shows that droplet ripening in concentration gradients exhibits
Droplet Ripening in Concentration Gradients
11
fundamental differences compared to classical phase-separating systems where droplet
positions are homogeneously distributed in space. The physics presented here could
be relevant for the control of emulsions in chemical engineering and biology. The
narrowing of droplet size distributions found in the presence of a regulator gradient
could be used to control droplet size in emulsions. It provides a physical mechanism
for the formation of almost mono-disperse emulsions. An example in biology where
an emulsion is controlled by concentration gradients is the C.elegans embryo [6, 7, 8].
In this system liquid-like cellular compartments, so called P granules, are positioned
toward the posterior side of the cell prior to asymmetric cell division by a protein
concentration gradient. An increasing number of membrane-less compartments with
liquid-like properties have been characterized [2, 21]. Their formation and positioning
could be a general scheme for the spatial organization of chemistry in living cells. In
our work we have identified the physical mechanisms of spatial segregation of droplets
by concentration gradients. The physics discussed here contributes to the behavior of
liquid-like compartments in living cells such as P granules. However, many aspects of
the dynamics of liquid-like compartments inside cells remain unexplored. In particular,
they consist of a large number of components and are chemically active. Emulsions in
the presence of chemical reactions driven away from equilibrium can give rise to novel
phenomena in phase separating systems such as the suppression of Ostwald ripening [22]
or the spontaneous division of liquid droplets [23]. Future questions could address how
nucleation, fusion, and droplet shape is changed by concentration gradients and how
non-equilibrium chemical reactions in droplets are affected by concentration gradients.
Acknowledgments
We would like to thank Shambaditya Saha and Anthony A. Hyman for stimulating
discussions.
References
[1] Brangwynne CP. Soft active aggregates: mechanics, dynamics and self-assembly of liquid-like
intracellular protein bodies. Soft Matter. 2011;7(7):3052–3059.
[2] Hyman AA, Weber CA, Julicher F. Liquid-liquid phase separation in biology. Annual review of
cell and developmental biology. 2014;30:39–58.
[3] Brangwynne CP, Tompa P, Pappu RV. Polymer physics of intracellular phase transitions. Nature
Physics. 2015;11(11):899–904.
[4] Cowan CR, Hyman AA. Asymmetric cell division in C. elegans: cortical polarity and spindle
positioning. Annu Rev Cell Dev Biol. 2004;20:427–453.
[5] Betschinger J, Knoblich JA. Dare to be different: asymmetric cell division in Drosophila, C.
elegans and vertebrates. Current biology. 2004;14(16):R674–R685.
[6] Brangwynne CP, Eckmann CR, Courson DS, Rybarska A, Hoege C, Gharakhani J, et al. Germline
P Granules Are Liquid Droplets That Localize by Controlled Dissolution/Condensation. Science.
2009;324(5935):1729–1732.
[7] Lee CF, Brangwynne CP, Gharakhani J, Hyman AA, Julicher F. Spatial Organization of the Cell
Cytoplasm by Position-Dependent Phase Separation. Phys Rev Lett. 2013 Aug;111:088101.
Droplet Ripening in Concentration Gradients
12
[8] Saha S, Weber CA, Nousch M, Adame-Arana O, Hoege C, Hein MY, et al. Polar Positioning
of Phase-Separated Liquid Compartments in Cells Regulated by an mRNA Competition
Mechanism. Cell. 2016;166(6):1572–1584.
[9] Lifshitz IM, Slyozov VV. The kinetics of precipitation from supersaturated solid solutions. Journal
of Physics and Chemistry of Solids. 1961;19(1?2):35 – 50.
[10] Wagner C. Theorie der Alterung von Niederschlagen durch Umlosen (Ostwald-Reifung). Berichte
der Bunsengesellschaft fur physikalische Chemie. 1961;65(7):581–591.
[11] Yao JH, Elder K, Guo H, Grant M. Theory and simulation of Ostwald ripening. Physical review
B. 1993;47(21):14110.
[12] Bray AJ. Theory of phase-ordering kinetics. Advances in Physics. 1994;43(3):357–459.
[13] Tenlen J, Molk J, London N, Page B, Priess J. MEX-5 asymmetry in one-cell C. elegans embryos
requires PAR-4- and PAR-1-dependent phosphorylation. Development. 2008;135(22):3665–3675.
[14] Griffin E, Odde D, Seydoux G. Regulation of the MEX-5 Gradient by a Spatially Segregated
Kinase/Phosphatase Cycle. Cell. 2011;146(6):955–968.
[15] Rubinstein M, Colby RH. Polymer physics. Oxford: OUP Oxford; 2003.
[16] We checked that a quench of the form Φout
eq (x) = Φ0[1 − s(x − L/2)] leads to qualitatively similar
results;.
[17] See Supplemental Material for videos and more information at http://;.
[18] The arrest phase is defined as the time interval during which N (t) decreases more slowly then
t−1/2 For Ostwald ripening (s = 0) N (t) ∼ t−1;.
[19] Clark MD, Kumar SK, Owen JS, Chan EM. Focusing Nanocrystal Size Distributions via
Production Control. Nano Letters. 2011 may;11(5):1976–1980.
[20] Vollmer J, Papke A, Rohloff M. Ripening and focusing of aggregate size distributions with overall
volume growth. Frontiers in Physics. 2014;2:18.
[21] Alberti S, Hyman AA. Are aberrant phase transitions a driver of cellular aging? BioEssays.
2016;38(10):959–968.
[22] Zwicker D, Hyman AA, Julicher F. Suppression of Ostwald ripening in active emulsions. Phys
Rev E. 2015 Jul;92:012317.
[23] Zwicker D, Seyboldt R, Weber CA, Hyman AA, Julicher F. Growth and division of active droplets
provides a model for protocells. Nature Physics. 2016;10.1038/nphys3984:1745–2481.
|
1601.01867 | 1 | 1601 | 2016-01-08T13:08:11 | On modelling of physical effects accompanying the propagation of action potentials in nerve fibres | [
"physics.bio-ph"
] | The recent theoretical and experimental studies have revealed many details of signal propagation in nervous systems. In this paper an attempt is made to unify various mathematical models which describe the signal propagation in nerve fibres. The analysis of existing single models permits to select the leading physiological effects. As a result, a more general mathematical model is described based on the coupling of action potentials with mechanical waves in a nerve fibre. The crucial issue is how to model coupling effects which are strongly linked to the ion currents through biomembranes. | physics.bio-ph | physics | On modelling of physical effects accompanying the propagation of action
potentials in nerve fibres
Juri Engelbrecht, Tanel Peets, Kert Tamm, Martin Laasmaa, Marko Vendelin
Centre for Nonlinear Studies, Institute of Cybernetics at TUT, Akadeemia tee 21, Tallinn 12618, Estonia
E-mail: [email protected], [email protected], [email protected], [email protected], [email protected]
Abstract. The recent theoretical and experimental studies have revealed many details of signal propagation in nervous systems.
In this paper an attempt is made to unify various mathematical models which describe the signal propagation in nerve fibres. The
analysis of existing single models permits to select the leading physiological effects. As a result, a more general mathematical
model is described based on the coupling of action potentials with mechanical waves in a nerve fibre. The crucial issue is how to
model coupling effects which are strongly linked to the ion currents through biomembranes.
Key words: nerve fibres, action potentials, mechanical waves, biomembranes.
1.
Introduction
The modelling of nerve pulse propagation is historically related to the analysis of action potentials
which have been measured using electrophysical techniques. The celebrated Hodgkin-Huxley (HH) model
[1] describes the propagation of asymmetric pulse in a nerve fibre which can be considered as a cylindrical
biomembrane filled with axoplasm. Due to ion currents (mostly Na and K) through the membrane this
pulse is able to propagate with a constant shape with an overshoot responsible for the refraction time.
Hodgkin and Huxley [1] have proposed kinetic equations for governing ion currents and have established
experimentally the needed parameters.
In order to grasp mathematically the essence of the nerve pulse
propagation, several simpler models have been proposed (FitzHugh-Nagumo [2]; Engelbrecht [3], etc). A
recent study [4] has summarised the present knowledge on mathematical modelling of action potentials
including neural networks.
In parallel to studies of action potentials, it has been shown experimentally
that mechanical displacements and heat transfer in axonal membrane accompany the electrical signal [5,
6]. In recent years these studies have been intensified thanks to the growing interest to the behaviour of
biomembranes which are the main building blocks of biology [7].
The starting point in modelling is the description of an electrical signal (action potential) which propa-
gates in axoplasm surrounded by a cylindrical biomembrane and is understood as a change in the potential
within the axoplasm. This cylindrical biomembrane is surrounded from outside by the intersticial fluid.
The action potential has a constant shape which is supported by ion currents through the biomembrane.
The well-known models for action potentials like HH, FitzHugh-Nagumo (FHN) or evolution equations
[3] include mechanisms of ion currents which according to the terminology of HH, are phenomenological
variables. From the viewpoint of thermodynamics, however, these can be considered as internal variables
[8].
Next, the processes in biomembranes must be understood. Much attention is paid to the fundamentals
of biomembranes which play an important role in many cellular processes including propagation of nerve
impulses (overview by Mueller and Tyler [7]). Two main problems can be distinguished: mechanisms of
opening ion channels and mechanisms of mechanical behaviour. The structure of biomembranes is nowa-
days well understood: they are made of phospholipids [7, 9] in the form of bilayers which can exist either in
a liquid phase or in a gel phase. Such a membrane is able to deform because of its elastic properties [10] in
the longitudinal and transverse direction as well as to undergo bending [7]. The studies involve explanation
of dynamical deformation of membranes [11].
The next important feature related to the action potentials is that it generates an axoplasmic pressure
pulse propagating synchronously with the action potential [12]. Such a pressure pulse will die out without
the constant support from an action potential as shown by Rvachev [12].
6
1
0
2
n
a
J
8
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
6
8
1
0
.
1
0
6
1
:
v
i
X
r
a
2
Finally, the closing link to the full description of the physiological phenomena in nerve fibres is the
impact of both the action potential and the pressure pulse on the membrane which should explain opening
the ion channels, the deformation of membranes, the generation of heat and accompanying optical effects
[11, 7, 13].
An important element of signal propagation in neural systems beside the propagation in fibres is the
transmittance of the signals across synaptic contacts to receptors (terminals). This transmittance is realised
by chemical and/or electrical signals [14, 15] and the electrical signalling may be interrelated with chemical
transmission [15]. Recent experiments have shown that a mechanical spike may also accompany the arrival
of the action potential to the terminal volume [16]. The role of a mechanical spike beside electrical and
chemical transmittance is not clear yet. However, as said by Bennet [17]: ‘the nervous system does its
operations in a number of different ways’.
It is a challenge to understand coupling between action potential and the accompanying effects together
with possible structural changes in biomembranes. Up to now, there is no theoretical consensus about the
general models which could describe all the measured phenomena within one framework. Clearly the struc-
tural properties of biomembranes [7] must be taken into account. Beside models based on electrophysics,
the models are proposed which prescribe the signal transfer to mechanics only [11, 13] or to protein ther-
modynamic structures [18]. Here in this paper we sum up the previous results (Section 2) and propose to
unite the various models into a more plausible description based on coupling of various effects (Section 3).
The brief discussion follows in Section 4.
2. Existing physical models and their parameters
2.1. Action potential
Both the axoplasm and the intersticial fluid contain ions of sodium (Na+) and potassium (K+) as well
as other ions [1, 19]. The relative concentration of ions create the transmembrane potential which at the
equilibrium value is in the range of -50 to -100 millivolts. If a stimulus applied to a nerve is below a certain
threshold value then the equilibrium value will be restored fast. If a stimulus is large enough (above the
threshold), then a stable action potential will be formed and it starts to propagate along the nerve fibre.
For a standard HH axon [19] with a diameter about 0.5 mm, the amplitude is about +50 millivolts and it
has a typical overshoot of about 20 millivolts. The whole duration of a pulse is about 4 -6 msec and the
propagation velocity is about 20 m/sec or more. The Hodgkin-Huxley (HH) model which describes the
propagation of such a pulse is the following [1]: (i) the basic telegraph equation where the inductance is
neglected (a diffusion-type equation):
∂ 2v
∂ x2 = crs
∂ v
∂t + rsIi
(1)
where v is the voltage, c - the capacitance, rs the resistance determined as 1/(πa2σ1), a is the radius of an
axon and σ1 - the conductivity and Ii is the ion current.
(ii) a phenomenological expression for the ion current:
Ii = gKn4(v− vR − vK) + gNam3h(v− vR − vNa) + gL(v− vR − vL) +Cm
∂ v
∂t ,
gK = 2πaGK,
gNa = 2πaGNa,
gL = 2πaGL.
(2)
(3)
Here GK and GNa are the maximum potassium and sodium conductances, respectively; GL is a constant
leakage conductance, Cm is the membrane capacitance per unit area, vR is the resting potential and vK,
vNa, vL are the corresponding equilibrium potentials. The phenomenological (hidden) variables n, m, and h
govern the ”turning on” and ”turning off” the membrane conductances.
(iii) kinetic equations for three phenomenological variables:
dn
dm
dt = αn(1− n)− βnn,
dt = αm(1− m)− βmm,
dt = αh(1− h)− βhh,
dh
where αi, βi are related to corresponding equilibrium values and relaxation times.
(iv) expressions for the parameters of the kinetic equations:
α = 0.01(10− v)[exp(10− v)/10− 1]−1,
βn = 0.125exp(−v/80),
αm = 0.01(25− v)[exp(25− v)/10− 1]−1,
βm = 4exp(−v/18),
αh = 0.07exp(−v/20),
βh = [exp(30− v)/10 + 1]−1.
3
(4a)
(4b)
(4c)
(5a)
(5b)
(5c)
(5d)
(5e)
(5f)
determined experimentally [1].
and reads [20]
The simplified FitzHugh-Nagumo (FHG) model takes into account only one phenomenological variable
∂ 3z
∂t∂ x2 =
∂ 2z
∂t2 + µ(1− z + εz2)
∂ z
∂t + z,
(6)
where µ and ε are parameters and z is the scaled voltage.
One can add to these traditional widely used models an evolution equation derived in [3]
∂ 2z1
∂ x∂ξ + f (z1)
∂ z1
∂ξ + g(z1) = 0,
(7)
where z1 is the scaled voltage and ξ is a moving frame ξ = c0t − x. Following the FHN model with one
phenomenological variable, f (z) is a quadratic function and g(z) - a linear function. Note that the moving
frame involves velocity c0 determined from the full telegraph equation. The final velocity c of a nerve pulse
is different from c0.
It must, however, be stressed that the ionic mechanisms are in fact more complicated than assumed even
in the HH model with three variables and include several pump, exchange and background currents [21].
The diversity of channels is reflected also in a wide range of shapes and patterns for nerve pulses [22].
The models described above are able to reflect the main features of measured action potentials: (i) the
existence of a threshold for an input; (ii) the all-or-non phenomenon for a pulse; (iii) the existence of an
asymmetric localised pulse with an overshoot; (iv) the existence of a refraction length; (v) the possible
annihilation of counter-propagating pulses [23, 20]. The propagation of an action potential is triggered by
an electrical excitation. A recent overview on processes in axons from the physiological viewpoint is given
in [14]. Despite of such clear and measurable characteristics of pulses, the governing equations of action
potentials can not describe other accompanying effects like mechanical displacements in the fibre walls or
heat transfers – see the recent criticism about the HH model by [24].
2.2. Biomembranes
Lipid membranes are not only a part of nerve fibres but they are important constituents of cells in much
wider context [7]. The behaviour of membranes under mechanical forces is experimentally observed. In the
4
equilibrium condition the membrane is in a liquid phase but under mechanical excitation the membrane un-
dergoes a structural change from the liquid phase to the gel phase [11, 13]. Consequently, its compressibility
will be changed, too. The measurements of compressibility are carried out by calorimetry [11, 25] and by
ultrasonic velocity measurements [25]. This is needed for building up the dynamical models for waves in
membranes like described in [26] where the energy density is given by a nonlinear expression:
e =
c2
0
ρ0
u2 +
p
3ρ0
u3 +
q
6ρ0
u4,
(8)
where c0 is the velocity of the longitudinal wave in a membrane, ρ0 is the density and u = ∆ρ is the amplitude
of the change in the membrane area while p and q are constants.
The structural changes in membranes under forcing are not only related to phase transformation, these
are also related to opening the channels (pores) for ions [7, 9] and for water molecules [10]. Such mechanosen-
sitive changes play an important role for ion currents and can explain the role of anesthetics which may
depend on the expansion modulus of membranes [12]. The opening of channels under mechanical stress
and electric field is demonstrated also by molecular dynamics simulations [9, 27, 28].
2.3. Dynamics of membranes
As explained by [7], biomembranes are able to resist pressure, tension, stretch and bending. The defor-
mation can be induced by mechanical or electrical impact [29, 30, 31]. The pressure pulse have been studied
theoretically [29] and experimentally [30]. The theoretical model of Griesbauer et al [29] for pulses in lipid
monolayers at the air-water interface is a second-order wave equation which is viscously coupled (the first
derivative) to the liquid (water) underneath. The existence of 2D sound waves along a lipid monolayer is
demonstrated by Shrivastava and Schneider [30]. The localized self-supporting pulses have been found hav-
ing a threshold amplitude and all-or-none nature. The need to take nonlinearity in compressibility together
with dispersive effects into account is stressed.
Such a nonlinear model with dispersion is proposed by Heimburg and Jackson [26], Andersen et al [13],
etc. The governing equation for a 1D pulse in a cylindrical biomembrane is proposed by using the energy
density (8) resulting in
(cid:20)(cid:0)c2
0 + pu + qu2(cid:1) ∂ u
(cid:21)
∂ x
− h
∂ 4u
∂ x4 ,
∂ 2u
∂t2 =
∂
∂ x
(9)
where u is the density change (cf. (8)) and h is an ad hoc constant. This is a Boussinesq-type equation where
nonlinearity is caused by the compressibility which has an impact on the velocity:
c2 = c2
0 + pu + qu2.
(10)
Due to the existence of nonlinearities and dispersion in Eq. (8), the solution of Eq. (8) may have a solitary
character [26]. In order to improve dispersive properties of such a model, it has been proposed to use more
realistic dispersion terms like it is done for modelling waves in rods [32]:
(cid:20)(cid:0)c2
0 + pu + qu2(cid:1) ∂ u
(cid:21)
∂ x
∂ 2u
∂t2 =
∂
∂ x
− h1
∂ 4u
∂ x4 + h2
∂ 4u
∂ x2∂t2
(11)
where h1, h2 are dispersive constants.
In order to compare these theoretical results with experiments [5] one has to understand that the Eqs.
(9), (10) describe longitudinal waves and the transverse displacements w measured by [5] can be found from
the derivative of the longitudinal profile [32, 33]:
w = −kr
∂ u
∂ x ,
(12)
5
where r is the radius of the axon and k is a constant. In the theory of rods k is the Poisson ratio. However, it
is unclear how fast the longitudinal waves in biomembranes decay without amplification.
The wave equations (9) and (11) density waves in biomembranes and are able to model localised pulses
for given nonlinearities and dispersive terms. As far as these localised pulses are of a solitary character, it
is proposed to call such models ‘ a soliton model’ for signals in nerve fibres [11, 13, 26, 34]. It should be
stressed that although such a signal corresponds to changes in a biomembrane from a liquid to a gel state
[7, 26], in its essence it is a longitudinal wave and the transverse displacement (swelling of a fibre) must be
calculated according to expression (12). It is shown that nonlinear equations (9) and (11) possess soliton-
type solution [26, 35, 32, 36]. Compared with the action potential, a soliton has not got an overshoot.
It is argued that higher density in a soliton should be accompanied by a dilated region due to the mass
conservation and this should prevent close sequence of pulses [37]. Intuitively it is understood but there
is no evidence about such a phenomenon in modelling. According to [26, 35], a soliton is formed by an
initial input [35] that is a pressure change. It is stressed that the mechanical wave (soliton) is associated
with a transient voltage change which affects the membrane potential difference [24] or, in other words ‘the
appearance of a voltage pulse is merely a consequence of the piezo-electric nature of the nerve membrane
which is partially charged and asymmetric’ [35]. However, this mechanism seemingly needs more detailed
inspection. A good side of model (9) is that it allows the collision of pulses [35] which is experimentally
observed in giant axons of the earthworm [37].
From a viewpoint of mathematical physics, eqs (9) and (11) are of the Boussinesq type [38] reflecting
nonlinear and dispersive effects. However, the nonlinearity in terms of u in wave equations is of a different
character from the usual ux type. The soliton solutions to such equations are found [26, 35, 36] and the emer-
gence of soliton trains described [32, 36]. In the emergence process the influence of a specific nonlinearity
is clearly seen – for given physical parameters [26] the smaller solitons move faster than the larger ones
[39, 40] . Such an effect differs from the behaviour of soliton trains in the classical Boussinesq equation
[41].
2.4. Pressure waves inside cylindrical biomembranes
The axoplasm within a nerve fibre is actually a gel consisting 87% of water held together by cytoskele-
ton [42]. In modelling such a complicated structure is taken as a pseudoplastic fluid [43] or as a viscous
compressible fluid [12, 33]. The possible intra-axonal transport of substances in the fluid is not taken into
account [44]. It means that a pressure wave can be described by Navier-Stokes equations and in the first
approximation by the 1D model while the biomembrane can be treated as an elastic tube. In this sense there
is a similarity between blood flow in aorta [45, 46, 47, etc.] and pressure waves in intersticial fluid.
The governing equation for 1D pressure waves is a momentum balance [48]. In terms of longitudinal
velocity v it reads:
where ρ is the density, p is the pressure, µ is the viscosity and F - the body force, and v is the velocity.
In terms of pressure in the 2D setting for waves in fluid surrounded by a cylindrical tube (shell) the
governing equations are [49]:
(cid:18)∂ v
(cid:19)
ρ
∂t + v
∂ v
∂ x
+
∂ p
∂ x
− µ
(cid:18)∂ 2 p
f
∂ 2 p
∂t2 = c2
∂ 2U
ρ
∂t2 +
∂ 2V
∂t2 +
ρ
∂ x2 +
∂ p
∂ x = 0,
∂ p
∂ r = 0,
∂ 2v
∂ x2 = F,
(cid:19)
∂ 2 p
∂ r2 +
1
r
∂ p
∂ r
,
(13)
(14a)
(14b)
(14c)
where x and r are cylindrical coordinates, c f is the velocity and U, V are longitudinal and transverse dis-
placements, respectively.
6
2.5. Coupling
The ideas of coupling and interrelation of processes related to nerve pulse propagation have been men-
tioned already long time ago [50, 23, etc.]. The recent studies have explained several mechanisms respon-
sible for this complicated physiological phenomenon. The voltage-induced changes in membrane tension
and mechanically sensitive ion channels are described in [7]. It has been shown how the electric field can
generate tension and bending in a membrane which is itself described by a fluid model [51]. Concerning
directly the nerve fibres, there is an understanding that an action potential in the nerve fibre generates an
axoplasmic pressure wave in the intersticial fluid which propagates synchronously with the action potential
[12, 33]. Both can in principle activate the ion channels in the surrounding biomembrane as a result of
mechano-electrical activation which plays a crucial role in transmitting the signals. However, this process is
not completely understood as well as the coupling of action potential and the pressure wave although many
elements of such processes are analysed (Heimburg and Jackson [26], Rvachev [12], El Hady and Machta
[33], etc) together with experimental evidence (Tasaki [5], Iwasa et al [6], Shrivastava and Schneider [30],
etc). It is mentioned in [52] that the channel effects usually attributed to action potential are indistinguishable
from effects caused by localised pulses in the biomembrane which undergoes phase changes [7]. One idea is
to describe the coupling by designating the potential energy for a whole system in the biomembrane and the
kinetic energy in the movement of the axoplasmic fluid [33]. Following this model, there are surface modes
in the biomembrane driven by changes in charges (electrical pulse) which generate compressive forces on
the membrane. The analysis is carried out by using Fourier transform. Rvachev [12] focuses on estimating
the velocities of an action potential and the pressure wave driven by it without making an assumption for
separate components responsible for kinetic and potential energy like in [33].
This is a fundamental question how to model coupling of all processes in a nerve fibre. Clearly a
simple combination of action potential models (like HH model) and the models of mechanical waves in
biomembranes is not possible. The signal transmittance in neural systems is a complex process with many
constituents and many coupling links reflecting the complexity of physiology [14, 53]. Based on the analysis
described above, we shall propose a coupled model based on following assumptions: (i) an action poten-
tial will generate a pressure wave; (ii) both electric field and mechanical impact can generate opening the
ion channels in the membrane; (iii) in the cylindrical biomembrane which surrounds the intersticial fluid,
mechanical waves are initiated synchronously with wave processes in the fluid; (iv) biomembrane can ex-
ist either in a fluid or in gel state. In the next Section these phenomena will be analysed within one fully
coupled model.
3. Fully coupled model
The key elements in modelling seem to be a biomembrane and ion channels which actually govern the
amplification of a nerve pulse (an action potential). The simulation of biomembranes up to now is not
possible with fully molecular setting [54]. That is why it is proposed to use continuum mechanics approach
and introduce forces for modelling possible coupling [54].
The simplest idea is to model nerve pulse by one of the existing action potential models (either by the
HH or by the simpler FHN model), the pressure wave in the intersticial fluid by using the Navier-Stokes
description and the waves in the cylindrical biomembrane by the improved Heimburg-Jackson model. We
base our modelling on the following ideas:
(i) electrical signals are the carriers of information [14] and trigger all the other processes;
(ii) the axoplasm in a fibre can be modelled as fluid where a pressure wave is generated due to electrical
signal; here, for example, the actin filaments in the axoplasm may influence the opening of channels in the
surrounding biomembrane but do not influence the generation of pressure wave in the fluid [42, 43, 33] ;
(iii) the biomembrane is able to deform (stretch, bending) under mechanical impact [29, 30];
(iv) the channels in biomembranes can be opened and closed under the influence of electrical signals as
well as of mechanical input; it means that tension of a membrane leads to the increase of transmembranal
ion flow and the intracellular actin filaments may influence the motions at the membrane [7, 10].
(v) there is strong experimental evidence on electrical or chemical transmittance of signals from one
neuron to another [17, 15].
The crucial problem to be solved is how to model the coupling between the action potential, pressure
wave, deformation of the biomembrane and opening/closing the ion channels. At this point we leave the
possibility of transfer the water molecules through the membrane aside.
7
Fig. 1. The flowchart of coupled models (left) and schemes of waves (right): AP - action potential PW - pressure wave in axoplasm,
LW - longitudinal wave in the biomembrane (BM), TW - transverse wave in the BM. Scales are arbitrary.
The flowchart of the coupled models is shown and the possible wave profiles are schematically depicted
in Fig. 1. The proposed set of coupled equations is the following:
Initial condition:
zt=0 = f (x),
where z is an electrical pulse. The action potential is governed by a FHN model:
(15)
(16)
(17)
(18)
where a1, a2 reflect electrical activation and b1(v), b2(v) - mechanical activation. The pressure wave in
axoplasm is governed by a Navier-Stokes model:
∂ z
∂t + z,
∂ 3z
∂t∂ x2 =
(cid:19)
∂ 2z
∂t2 + µ(1− (a1 + b1)z + (a2 + b2)z2)
(cid:18)∂ v
(cid:20)(cid:0)c2
0 + pu + qu2(cid:1) ∂ u
= −∂ p
∂ x + µν
(cid:21)
∂ 2v
∂ x2 + F1(z),
∂t + v
∂ v
∂ x
ρ
∂ 2u
∂t2 =
∂
∂ x
− h1
∂ 4u
∂ x4 + h2
∂ 4u
∂ x2∂t2 + F2(z,v),
∂ x
where v is the velocity, ρ the density, µν - the viscosity and F1(z) is a force from the action potential. In the
biomembrane, the longitudinal wave is governed by:
where the notations follow Eqs. (8), (9) and F2(z,v) is a force from the process in the axoplasm. In the
biomembrane the transverse wave is governed by
w = −kr
∂ u
∂ x ,
(19)
where r = a is the axon radius and k is the coefficient which in the theory of rods is the Poisson ratio
but in the present case dues to the complicated structure of a biomembrane, needs to be determined from
experiments.
8
4. Discussion
The coupled model (16) – (19) is a set of partial differential equations with coupling forces F1(z),
F2(z,v). The process is generated by an electrical impulse (15). Note that according to the soliton model
[26] the starting point is a mechanical stimulus which can change the electrical potential and the charge of
the membrane [24]. Although all the basic single models are well analysed and understood, the coupling
forces lead to many open questions. Starting with an input for the action potential, due to the coupling
(reciprocity) it should also be possible to model the influence of a mechanical wave in a biomembrane
to the action potential. Especially important is how electrical and/or mechanical waves either separately or
jointly influence the opening of ion channels [7]. Actually one could model this influence either by adjusting
nonlinear parameters ai , bi in Eq. (16) or by an additional external force F3(u,v) which acts on the initial
system. In the case of the FHN model (16) this assumption means that ai +bi = const and an additional term
∂
∂t F3(u,v), k0 = const appears on the r.h.s. of Eq. (16). Note that Eqs. (16), (18) are conservative and
like k0
Eq. (17) includes possible dissipation due to viscosity. The possible heat changes are not included into the
model but it is possible to account them by using the concept of internal variables [55]. Another important
question is to specify the physiological constants which certainly can vary over a large scale [14]. It must
be noted that the proposed mathematical model is a robust one which links the basic wave phenomena.
The action potential within this fully coupled model is described by the FHN model, which takes only one
ion current into account. Certainly, the more detailed HH model can be used. It is not clear yet how the
overshoot of an action potential influences the mechanical wave in a biomembrane. Even more, there are
several specific phenomena like the variation of ion channel densities, the extracellular accumulation of of
ions, the movement of mitochondria, etc which all affect the action potential propagation [56].
The crucial problem in modelling is to determine the coupling forces between the physical processes
(waves). It must be noted that the present model describes the coupling of variables at a given time moment
but the coupling could also be time-dependent taking into account the history of the process. In this case the
forces should be described by convulution-type integrals. Here the ‘history’ in some sense is related to the
overshoot of an action potential which keeps the sequence of pulses separated by the diffraction length. It
is proposed that as far as the mean area of density of the biomembrane stays constant then the compressed
regions (solitary wave) should be followed by regions of negative density change [52] and this explains the
minimum distance between solitary waves.
Another important problem is related to thermodynamics of the full process [13]. In order to deform
the membrane, a certain amount of work must be done which can be related to the free energy density like
Eq. (8). The general theory based on continuum mechanics derived for biomembranes is given in [58, 59],
in terms of Canham-Helfrich free energy in [60] and more specifically related to nerve pulse propagation in
[13, 26, 52].It might be useful to explain the work done for deformation directly to energy function like it is
done for cardiac mechanoenergetics [61] because this might give new insight to the phase transition in the
biomembrane.
Clearly, future studies are needed especially for clarifying the coupling processes. The recording tech-
niques [15, 14, 57] developed for studying axon physiology are promising for detecting the characteristics
of processes with great accuracy. This is needed not only for determining the parameters but also for under-
standing the physiological processes. For example, in most studies it is accepted that that an action potential
is generated by a synaptic input. In recent studies it is reported that an action potential can be initiated also
from the distal axonal region [57]. Due to possible coupling at various levels, the model described in Section
3, might allow such a phenomenon. The authors hope that the modelling principles described above could
help understanding this complicated physiological process.
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 33-7, IUT 33-24)
9
REFERENCES
1. A. L. Hodgkin and A. F. Huxley. A quantitative description of membrane current and its application to conduction and
excitation in nerve. J. Physiol., 117(4):500–544, 1952.
2. R. Fitzhugh. Impulses and physiological states in theoretical models of nerve membrane. Biophys. J., 1(6):445–466, 1961.
3. J. Engelbrecht. On theory of pulse transmission in a nerve fibre. Proc. R. Soc. London, 375(1761):195–209, 1981.
4. P. C. Bresslof. Waves in Neural Media. Springer, 2014.
5. I. Tasaki. A macromolecular approach to excitation phenomena: mechanical and thermal changes in nerve during excitation.
Physiol. Chem. Phys. Med. NMR, 20:251–268, 1988.
6. K. Iwasa, I. Tasaki, and R. Gibbons. Swelling of nerve fibers associated with action potentials. Science, 210(4467):338–339,
1980.
7. J. K. Mueller and W. J. Tyler. A quantitative overview of biophysical forces impinging on neural function. Phys. Biol.,
11(5):051001, 2014.
8. G. A. Maugin and J. Engelbrecht. A thermodynamical viewpoint on nerve pulse dynamics. J. Non-Equilibrium Thermodyn.,
19(1), 1994.
9. D. P. Tieleman, H. Leontiadou, A. E. Mark, and S.-J. Marrink. Simulation of pore formation in lipid bilayers by mechanical
stress and electric fields. J. Am. Chem. Soc., 125(21):6382–3, 2003.
10. H. Barz, A. Schreiber, and U. Barz. Impulses and pressure waves cause excitement and conduction in the nervous system.
Med. Hypotheses, 81(5):768–72, 2013.
11. T. Heimburg. Mechanical aspects of membrane thermodynamics. Estimation of the mechanical properties of lipid membranes
close to the chain melting transition from calorimetry. Biochim. Biophys. Acta, 1415(1):147–62, 1998.
12. M. M. Rvachev. On axoplasmic pressure waves and their possible role in nerve impulse propagation. Biophys. Rev. Lett.,
5(2):73–88, 2010.
13. S. S. L. Andersen, A. D. Jackson, and T. Heimburg. Towards a thermodynamic theory of nerve pulse propagation. Prog.
Neurobiol., 88(2):104–13, 2009.
14. D. Debanne, E. Campanac, A. Bialowas, E. Carlier, and G. Alcaraz. Axon physiology. Physiol. Rev., 91(2):555–602, 2011.
15. S. G. Hormuzdi, M. A. Filippov, G. Mitropoulou, H. Monyer, and R. Bruzzone. Electrical synapses: A dynamic signaling
system that shapes the activity of neuronal networks. Biochim. Biophys. Acta - Biomembr., 1662(1-2):113–137, 2004.
16. G. H. Kim, P. Kosterin, a. L. Obaid, and B. M. Salzberg. A mechanical spike accompanies the action potential in Mammalian
nerve terminals. Biophys. J., 92(9):3122–9, 2007.
17. M. V. L. Bennett. Electrical synapses, a personal perspective (or history). Brain Res. Rev., 32(1):16–28, 2000.
18. Q. Zhao. A molecular and biophysical model of the biosignal. Quantum Matter, 2(1):9–16, 2013.
19. A. Scott. Nonlinear Science. Emergence & Dynamics of Coherent Structures. Oxford University Press, 1999.
20. J. Nagumo, S. Arimoto, and S. Yoshizawa. An active pulse transmission line simulating nerve axon. Proc. IRE, 50(10):2061–
2070, 1962.
21. M. Courtemanche, R. J. Ramirez, and S. Nattel. Ionic mechanisms underlying human atrial action potential properties : insights
from a mathematical model. Am. J. Physiol., 275(1):H301–H321, 1998.
22. B. P. Bean. The action potential in mammalian central neurons. Nat. Rev. Neurosci., 8(6):451–65, 2007.
23. A. L. Hodgkin and A. F. Huxley. Resting and action potentials in single nerve fibres. J. Physiol., 104:176–195, 1945.
24. R. Appali, U. Van Rienen, and T. Heimburg. A comparison of the Hodgkin-Huxley model and the soliton theory for the action
potential in nerves. In A. Iglic, ed., Adv. Planar Lipid Bilayers Liposomes, Vol. 16, 275–299. Academic Press, 2012.
25. W. Schrader, H. Ebel, P. Grabitz, E. Hanke, T. Heimburg, M. Hoeckel, M. Kahle, F. Wente, and U. Kaatze. Compressibility of
lipid mixtures studied by calorimetry and ultrasonic velocity measurements. J. Phys. Chem. B, 106(25):6581–6586, 2002.
26. T. Heimburg and A. D. Jackson. On soliton propagation in biomembranes and nerves. Proc. Natl. Acad. Sci. USA,
102(28):9790–5, 2005.
27. J. C. Shillcock and R. Lipowsky. Equilibrium structure and lateral stress distribution of amphiphilic bilayers from dissipative
particle dynamics simulations. J. Chem. Phys., 117(10):5048–5061, 2002.
28. R. A. Bockmann, B. L. de Groot, S. Kakorin, E. Neumann, and H. Grubmuller. Kinetics, statistics, and energetics of lipid
membrane electroporation studied by molecular dynamics simulations. Biophys. J., 95(4):1837–50, 2008.
29. J. Griesbauer, S. Bossinger, A. Wixforth, and M. F. Schneider. Propagation of 2D pressure pulses in lipid monolayers and its
possible implications for biology. Phys. Rev. Lett., 108(19):198103, 2012.
30. S. Shrivastava and M. F. Schneider. Evidence for two-dimensional solitary sound waves in a lipid controlled interface and its
implications for biological signalling. J. R. Soc. Interface, 11(97):20140098, 2014.
31. A. G. Petrov. Flexoelectricity of model and living membranes. Biochim. Biophys. Acta, 1561:1–25, 2001.
32. J. Engelbrecht, K. Tamm, and T. Peets. On mathematical modelling of solitary pulses in cylindrical biomembranes. Biomech.
Model. Mechanobiol., 14:159–167, 2015.
33. A. El Hady and B. B. Machta. Mechanical surface waves accompany action potential propagation. Nat. Commun., 6:6697,
2015.
34. R. Appali, B. Lautrup, T. Heimburg, and U. Van Rienen. Soliton collision in biomembranes and nerves-a stability study. J.
10
Math. Ind., 16:205–212, 2012.
35. B. Lautrup, R. Appali, A. D. Jackson, and T. Heimburg. The stability of solitons in biomembranes and nerves. Eur. Phys. J. E.
Soft Matter, 34(6):1–9, 2011.
36. T. Peets, K. Tamm, and J. Engelbrecht. Numerical investigation of mechanical waves in biomembranes. In S. Elgeti and J.-W.
Simon, editors, Conf. Proc. YIC GACM 2015 3rd ECCOMAS Young Investig. Conf. 6th GACM Colloquium, July 20-23,
2015, Aachen, Ger., pages 1–4, 2015.
37. A. Gonzalez-Perez, R. Budvytyte, L. D. Mosgaard, S. Nissen, and T. Heimburg. Penetration of action potentials during
collision in the median and lateral giant axons of invertebrates. Phys. Rev. X, 4(3):031047, 2014.
38. C. I. Christov, G. A. Maugin, and A. V. Porubov. On Boussinesq’s paradigm in nonlinear wave propagation. Comptes Rendus
M´ecanique, 335(9-10):521–535, 2007.
39. K. Tamm and T. Peets. On solitary waves in case of amplitude-dependent nonlinearity. Chaos, Solitons & Fractals, 73:108–
114, 2015.
40. T. Peets and K. Tamm. On mechanical aspects of nerve pulse propagation and the Boussinesq paradigm. Proc. Est. Acad. Sci.,
64(3S):331–337, 2015.
41. J. Engelbrecht, A. Salupere, and K. Tamm. Waves in microstructured solids and the Boussinesq paradigm. Wave Motion,
48(8):717–726, 2011.
42. D. S. Gilbert. Axoplasm architecture and physical properties as seen in the Myxicola giant axon. J. Physiol., 253:257–301,
1975.
43. R. Biondi, M. Levy, and P. Weiss. An engineering study of the peristaltic drive of axonal flow. Proc Natl Acad Sci U S A,
69(7):1732–1736, 1972.
44. P. A. Weiss. Neuronal dynamics and axonal flow: axonal peristalsis. Proc. Natl. Acad. Sci., 69(5):1309–1312, 1972.
45. T. Pedley. The Fluid Mechanics of Large Blood Vessels. Cambridge University Press, 1980.
46. T. Moodie, D. Barclay, and R. Tait. A boundary value problem for fluid-filled viscoelastic tubes. Math. Model., 4:195–207,
1983.
47. D. Bessems, M. Rutten, and F. Van De Vosse. A wave propagation model of blood flow in large vessels using an approximate
velocity profile function. J. Fluid Mech., 580:145–168, 2007.
48. J. Tritton. Physical Fluid Dynamics. Oxford Sci. Publ., 1988.
49. T. Lin and G. Morgan. Wave propagation through fluid contained in a cylindrical elastic shell. J. Acoust. Soc. Am., 28(6):1165–
1176, 1956.
50. E. Wilke. On the problem of nerve excitation in the light of the theory of waves. Pflugers Arch., 144:35–38, 1912.
51. D. Lacoste, M. C. Lagomarsino, and J. F. Joanny. Fluctuations of a driven membrane in an electrolyte. Europhys. Lett.,
77(1):18006, 2007.
52. E. V. Vargas, A. Ludu, R. Hustert, P. Gumrich, A. D. Jackson, and T. Heimburg. Periodic solutions and refractory periods in
the soliton theory for nerves and the locust femoral nerve. Biophys. Chem., 153(2-3):159–67, 2011.
53. F. Contreras, H. Cervantes, M. Aguero, and M. d. L. Najera. Classic and non-classic soliton like structures for traveling nerve
pulses. Int. J. Mod. Nonlinear Theory Appl., 2:7–13, 2013.
54. F. L. H. Brown. Elastic modeling of biomembranes and lipid bilayers. Annu. Rev. Phys. Chem., 59:685–712, 2008.
55. A. Berezovski, J. Engelbrecht, and G. A. Maugin. Thermoelasticity with dual internal variables. J. Therm. Stress., 34(5-
6):413–430, 2011.
56. D. J. Bakkum, U. Frey, M. Radivojevic, T. L. Russell, J. Muller, M. Fiscella, H. Takahashi, and A. Hierlemann. Tracking
axonal action potential propagation on a high-density microelectrode array across hundreds of sites. Nat. Commun.,
4:2181, 2013.
57. T. Sasaki. The axon as a unique computational unit in neurons. Neurosci. Res., 75(2):83–88, 2013.
58. M.A. Lomholt and L. Miao. Descriptions of membrane mechanics from microscopic and effective two-dimensional perspec-
tives. J. Phys. A: Math. Gen., 39:10323–10354, 2006.
59. L. Deseri and G. Zurlo. The stretching elasticity of biomembranes determines their line tension and bending rigidity. Biomech.
Model. Mechanobiol., 12:1233–1242, 2013.
60. T. Chou.
Physics
of
cellular materials:
biomembranes
[lecture
notes].
Retrieved
from
http://faculty.biomath.ucla.edu/tchou/pdffiles/lecture3.pdf, 2002.
61. M. Kalda, P. Peterson, J. Engelbrecht and M. Vendelin. A cross-bridge model describing the mechanoenergetics of actomyosin
interaction. In G.A. Holzapfel and E. Kuhl, editors, Computer Models in Biomechanics, Springer Netherlands, pages
91–102, 2013.
|
1201.6270 | 1 | 1201 | 2012-01-30T16:18:48 | Is the immune network a complex network? | [
"physics.bio-ph",
"q-bio.MN"
] | Some years ago a cellular automata model was proposed to describe the evolution of the immune repertoire of B cells and antibodies based on Jerne's immune network theory and shape-space formalism. Here we investigate if the networks generated by this model in the different regimes can be classified as complex networks. We have found that in the chaotic regime the network has random characteristics with large, constant values of clustering coefficients, while in the ordered phase, the degree distribution of the network is exponential and the clustering coefficient exhibits power law behavior. In the transition region we observed a mixed behavior (random-like and exponential) of the degree distribution as opposed to the scale-free behavior reported for other biological networks. Randomness and low connectivity in the active sites allow for rapid changes in the connectivity distribution of the immune network in order to include and/or discard information and generate a dynamic memory. However it is the availability of the low concentration nodes to change rapidly without driving the system to pathological states that allow the generation of dynamic memory and consequently a reproduction of immune system behavior in mice. Although the overall behavior of degree correlation is positive, there is an interplay between assortative and disassortative mixing in the stable and transition regions regulated by a threshold value of the node degree, which achieves a maximum value on the transition region and becomes totally assortative in the chaotic regime. | physics.bio-ph | physics | Is the immune network a complex network?
Hallan Souza-e-Silva∗ and Rita Maria Zorzenon dos Santos†
Departamento de F´ısica, Centro de Ciencias Exatas e da Natureza,
Universidade Federal de Pernambuco, Recife, PE, Brasil
Some years ago a cellular automata model was proposed to describe the evolution of the immune
repertoire of B cells and antibodies based on Jerne's immune network theory and shape-space
formalism.
Interactions among different B-cell clones, which may be found in low, intermediate
and high concentrations, occur depending on the complementarities of their characteristic proteins
and are regulated by activation and suppression mechanisms. Depending on the region of the
parameter space, the model exhibits either stable (ordered) or chaotic behavior, but it is in the
transition region of the parameter space between both regimes where we obtain the complex behavior
of a self-regulated network, much like Jernes idea of the immune network.
In this region, the
network maintains the memory of the large perturbations, which simulate antigen presentations
and reproduce immunization and ageing experiments performed with mice. Here we investigate if
the networks generated by this model in the different regimes can be classified as complex networks.
We have found that in the chaotic regime the network has random characteristics with large, constant
values of clustering coefficients, while in the ordered phase, the degree distribution of the network
is exponential and the clustering coefficient exhibits power law behavior. In the transition region
we observed a mixed behavior (random-like and exponential) of the degree distribution as opposed
to the scale-free behavior reported for other biological networks. Randomness and low connectivity
in the active sites allow for rapid changes in the connectivity distribution of the immune network
in order to include and/or discard information and generate a dynamic memory. However it is
the availability of the low concentration nodes to change rapidly without driving the system to
pathological states that allow the generation of dynamic memory and consequently a reproduction
of immune system behavior in mice. Although the overall behavior of degree correlation is positive,
there is an interplay between assortative and disassortative mixing in the stable and transition
regions regulated by a threshold value of the node degree, which achieves a maximum value on the
transition region and becomes totally assortative in the chaotic regime.
2
1
0
2
n
a
J
0
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
0
7
2
6
.
1
0
2
1
:
v
i
X
r
a
∗ [email protected], Permanent Address: Departamento de F´ısica, Centro de Ciencias Exatas, Universidade Federal de Vi¸cosa, Vi¸cosa,
MG, Brasil
† [email protected], corresponding author
I.
INTRODUCTION
2
The main task of the immune system is to protect the integrity and identity of the body against any harm. The
main cells of the immune system are macrophages and lymphocytes, the former being mostly responsible for innate
immune responses and lymphocytes, which are responsible for adaptive or cell-mediated immune responses.
All immune cells carry a large number of molecular receptors (proteins) on the surface and the immune system
works based on pattern recognition. There are two main classes of lymphocytes: T and B cells. While T cells are
involved in signaling and functional activities for the majority of the pathogens, B cells are mainly responsible for
the production of antibodies, which in general, function as markers for the pathogens to be phagocyted and improve
the efficiency of the immune response. The antibodies produced by any B-cell population are copies of its molecular
receptors and according to the clonal selection theory [1] proposed by Burnet in 1959, the antigen (or its binding
sites), by pattern recognition chooses the B-cell clones (population of B cell and antibodies) that will proliferate.
According to estimates lymphocytes carry the order of 105 molecular receptors on its surface and the human immune
system is able to express the order of 1011 different receptors during its lifetime. Such large numbers allow for the
recognition of any antigen presented to the immune system and for the completeness of the immune repertoire. If
the repertoire is complete we should expect the elements of the immune system to recognize and be recognized by
other elements; the same mechanism of recognition should also work for both antibody-antigen and antibody-antibody
reactions. In 1974, Jerne [2], taking these ideas into account, suggested that when the antigen is presented to the
organism it will activate a set of B-cell clones and the production of specific antibodies (complementary to the antigen
binding sites) that in turn would activate the receptors of other clones, and so on. Due to the interplay of activation
and suppression mechanisms, the reaction chain would be finite, preventing the percolation of information through
the entire system. This kind of dynamics generates a multi-connected network of cell populations that regulates the
immune response, and at any time reflects the dynamic memory of the system regarding its previous history in terms
of antigen presentation. In other words, the immune response to different pathogens (virus, bacteria, etc) is regulated
by dynamics involving the complementary molecular receptors of the different B-cell clones that create a dynamic
memory, which includes new and already existing information concerning the previous immunization process. This
memory allows for a fast response to new presentations of previously seen antigens.
Since its proposal, little evidence has been put forward to support the existence of the immune network theory
[3 -- 5], but research has suggested that if the network exists, then only 20% of the lymphocytes will be activated, while
the rest of the clones will form a pool of immunocompetent lymphocytes, which are able to recognize any antigen.
In 1992, Stauffer and Weisbush [6] proposed a cellular automata model to describe the immune network based on
Jerne's ideas, shape-space formalism and a previous model introduced by De Boer and Perelson [7]. This discrete
model was later modified by Zorzenon dos Santos and Bernardes [8] and hereafter will be referred to as the BSP
model. Using shape-space formalism, the model allows for the simulation of the large immune repertoire and the
complementary interactions between B-cell clones. Each B-cell population is associated to a three-state automatum
representing low, intermediate and high concentrations, and the interactions amongst the different populations are
described by activation and suppression mechanisms. The dynamics of the model leads to stable (ordered) or chaotic
behaviors that correspond to pathological rather than regular behaviors of the immune system. However, in the
transition region between the two regimes [9] during very long periods, the model describes aggregation-disaggregation
dynamics with clusters splitting and fusing throughout time, as a multi-connected network of populations [10]. This
network exhibits self-regulation and despite changes within the active populations, only 10-20% of the populations
remain active. This multi-connected network subjected to multiple perturbations was used with mice to simulate and
reproduce the behavior of their immune systems under multiple antigen presentations and the effects of ageing on
generating these responses [11].
The dynamic behavior of the BSP model on the transition region was further investigated regarding different aspects.
It has especially been observed that the memory of the system is dynamically allocated in order to incorporate new
information regarding the history of the system without losing previously acquired important information [12]. As
the system becomes older, so less information is incorporated, which is a behavior associated to the loss of plasticity,
as observed in other physical systems [13].
Although there have been many studies on the parameter space and dynamics of the BSP model, it has not yet
been characterized from the viewpoint of a complex network. Since it reproduces the behavior of real immune systems
it is interesting to investigate its topological properties using graph theory formalism (see [14] and references therein)
adopted in the characterization of complex networks found in nature [14, 15]. From previous studies on the dynamic
behavior of the immune network it becomes clear that the regulation of the dynamics leading to stable, complex or
chaotic behaviors emerges from the interplay between subnetworks of populations with low, intermediate and high
activations. Our aim in this work is to investigate whether the networks generated by the BSP model in the different
regimes (stable, transitional and chaotic) may be classified as complex networks. Therefore, we have focused on
studying and characterizing the properties of these subnetworks for each phase with regard to degree distribution, the
3
behavior of clustering coefficients, the existence of hierarchical structures and the correlations within the neighborhood
(assortative or disassortative mixing behaviors) of the sites.
In what follows, in section II, we introduce the SW cellular automata model proposed to describe the immune
network, in section III we undertake a short review of the main concepts involved in the study of the above-mentioned
properties, in section IV we present and discuss the results obtained and in section V we make our final remarks.
II. THE CELLULAR AUTOMATA MODEL
In the model introduced by Stauffer and Weisbuch [6] shape-space formalism is adopted to describe the repertoire
of B-cell clones and the interaction between different clones [16]. Each clone is represented by its molecular receptor,
which corresponds to a point in a d-dimensional space. Thus, each receptor is characterized by d properties cor-
responding to different characteristics, as for instance, the number of nucleotides, charge, hydrophobicity, etc. The
nearest clone neighbors differ only by one property and the lock-key interactions occurring among populations with
complementary molecular receptors is described by a non-local rule, where an on-site clone ~r is influenced by those
located at −~r and its nearest neighbors (representing slightly defective interactions). According to estimates, if the
shape-space notion is relevant in a continuous mathematical approach, we obtain d ≥ 5 [16] but if it is relevant in a
discrete approach, then it is d ≥ 2 [10]. The population associated to each molecular receptor ~r is represented by a
three-state automaton describing its concentration at any given time: low (B(~r, t) = 0), intermediate (B(~r, t) = 1)
and high (B(~r, t) = 2). The influence on the population of site ~r caused by its complementary populations is described
by the field h(~r, t):
h(~r, t) = X
~r′ǫ−(~r+δ~r)
B(~r, t)
(1)
where for each ~r the sum runs over the complementary shape −~r and its nearest neighbors. Due to the finite number
of B-cell population states, the maximum value of the field h(~r, t) is hmax = 2(2d + 1). The rules describing the
changes of B-cell populations due to their interactions with other populations are based on an activation window,
which is inspired by a log-bell-shaped proliferation function associated to the receptor cross-linking involved in B-cell
activation [6, 8, 10]. In this window there is a minimum field necessary to activate the proliferation of the receptor
populations (θ1), but there is also an upper limit for activation since for high doses of activation (greater than θ2) the
proliferation is suppressed. The updating rules are summarized as follow:
B(~r, t + 1) =
B(~r, t) + 1 if
B(~r, t) − 1
θ1 ≤ hi(t) ≤ θ2
otherwise
(2)
but no change is made if it would lead to B = −1 or B = 3. At each time step t we define the densities of sites in
state i as Bi(t), where i= 0, 1 and 2 correspond to low, intermediate and high concentrations, respectively.
The initial configurations are randomly generated depending on the parameter x that controls the initial concen-
trations: B1(0) = B2(0) =x/2, while the remaining Ld(1 − x) sites are initiated with low concentrations.
This model exhibits stable and chaotic regimes for d ≥ 2 [8], depending on the activation threshold (θ1) and the
width of the activation window (θ2 − θ1). These different dynamic behaviors are separated by a transition region
where the model exhibits complex behavior during cycles of very long periods [9] in which clusters of different B-cell
populations split and fuse [10] and the activated populations behave like a multi-connected network. In 1998 [11]
this multi-connected network was used to reproduce the results of experiments performed with mice and showed
refractory behavior under multiple antigen presentations. Antigen presentation is simulated by the introduction of
perturbations flipping the state of low concentration populations to those of high concentration, and computing the
changes on the multi-connected network with respect to the initial state before the introduction of the perturbation.
The same study has shown that the aging effects observed in the immune responses of mice can be explained by
the loss of plasticity and the ability to bring about new changes in order to include new information in the system,
i.e., the older the system, the more rapidly it saturates and the less intense is its response. It has also been found
[17] that there is a characteristic cluster size associated to the loss of plasticity and a power law distribution for
the permanence times (the time interval that each population remains activated or belongs to the multi-connected
network). As expected, the absence of scale indicates that there is no typical permanence time, a feature that
may be understood since the memory of the system is allocated dynamically at each time step depending on the
interactions among the populations. Since the aging effects observed in the multi-connected network have similarities
4
with physical glassy systems, auto-correlation functions were obtained [13]. As previously mentioned, the system
with no perturbation is driven to a long-limit cycle-attractor after a long transient time. When subjected to small,
random perturbations however, the very notion of a transient becomes fuzzy and the results in Ref. [13] show that the
system is deflected from its attractor by small perturbations after 103 − 104 time steps. Therefore, small perturbations
will cause the system to change attractors from time to time due to their cumulative effects, and is reflected in the
decreasing auto-correlation functions. Large perturbations would not accelerate the de-correlation process. In fact,
large perturbations lead to a much weaker (slower) de-correlation, since they are always produced on the same sites in
shape-space to simulate antigen presentations. Small perturbations can be more easily absorbed by the system than
larger ones since they involve only local changes. The various studies performed until now [10, 12, 13] to understand
the dynamics of the multi-connected network have suggested that different behaviors (stable, complex and chaotic)
emerge from the interplay between the subnetworks of low, intermediate and high concentrations due to the activation
of suppression mechanisms. In this work, we have investigated the properties of these inter-dependent sub networks
from the viewpoint of complex networks using graph theory formalism.
III. THE COMPLEX NETWORK PROPERTIES INVESTIGATED
In what follows we will consider the multi-connected network and the subnetworks as graphs composed of nodes
and links. The key quantity on the study of the topological properties of graphs and complex networks is the degree
distribution of the nodes. The connectivity or degree associated to the node i is defined by the number Ki of links
that connects node i to other nodes on the graph. In other words,
Ki = X
j
aij
(3)
where the sum runs over all nodes j not equal to i and aij corresponds to connectivity or adjacent matrix elements. If
there is a link between i and j, aij = 1, otherwise aij = 0. The degree distribution obtained for the network indicate
the topological class of universality (regular, random or scale-free) to which the complex network belongs and whether
there is an existence hidden variables that may deviate its behavior from one of the main universality classes.
Due to the complementary interaction that regulates the interactions between the B-cell populations, another
quantity of interest in the study of the multi-connected network is the clustering coefficient of a given node, which is
a measure of the degree to which nodes in a graph tend to cluster together. The clustering coefficient of a node i (Ci)
is defined as:
Ci = 2
N C
Ki(Ki − 1)
(4)
where N C is the number of connections between the neighbors of site i. The average of all nodes gives the network
clustering coefficient C. While Ci is a local property, C is global. The global clustering coefficient gives an overall
indication of the clustering in the network, whereas the local coefficient describes the embedding of single nodes.
The behavior of C(K) as a function of K may indicate the presence of hierarchical structures in the network when it
exhibits power-law behavior [18]. Hierarchical structures go beyond simple clustering by including the simultaneous
organization of all scales in a network. Such structures are represented by trees or dendrograms, in which closely
related pairs of vertices (nearest neighbors) have the lowest common ancestors than more distantly related pairs.
In the case of the immune network we should expect to find some hierarchical structures, since it is built through
complementary interactions using shape-space formalism, where similar shapes should have common ancestors.
To complete the characterization of the subnetworks we investigate the existence of assortative or disassortative
mixing on the different subnetworks, a measurement that will define if the highly-connected nodes in each subnetwork
tend to have more connections with other highly-connected nodes or with nodes with a lower number of connections,
respectively. In order to investigate the assortativity of the network we calculate the degree-correlation of the com-
plex network (Knn) [19]. When there is a tendency of the nodes to connect to other nodes with a similar degree,
the correlation is assortative, while when there is a prevalence of links between nodes with dissimilar degrees, the
correlations are disassortative. The average degree between nearest neighbors Knn [19] is defined as:
Knn(K) =
1
NK X
i/Ki=k
Knn(i)
(5)
5
The sum runs over all nodes with degree K, NK is the number of nodes of degree K and Knn(i) is the average
nearest neighbors degree of vertex i:
Knn(i) =
1
Ki X
j
Kj
(6)
where the sum runs over all nearest neighbors of node i. When Knn is the constant, the degrees of neighboring nodes
are uncorrelated.
Another quantity of interest is the mean shortest path (L) that corresponds to the average number of connections
between any two nodes i and j calculated over all pairs of nodes. For regular hypercubic lattices in d-dimension
L ∼ N 1/d and in the case of random graphs L grows logarithmically with the number N of nodes (L ∼ logN . The
small world effect corresponds to the case in which any pair of nodes is connected by the shortest path distance [19].
By investigating these properties it was possible to better understand the dynamics of the immune network from a
topological viewpoint and the interdependence of the three subnetworks.
IV. RESULTS
All the results reported in this section were obtained using the following parameters: d = 3, L = 50, N = Ld
sites, θ1 = 1/3 × hmax and θ2 = 2/3 × hmax, where hmax = 2(2d + 1) and they are representative of the behaviors
obtained on the different regions analyzed. Thus, when varying the parameters we should expect variations in the
localization and size of these regions in the parameter space, but the overall behavior of the quantities analyzed here
in the different regions would be the same. For the sake of clarity we will refer to the subpopulations or subnetworks
of different states of activation as: B0 for low, B1 for intermediate and B2 for high. By subnetwork we mean the
network formed by all nodes in the same state. In the present study, for the maximum number of neighbors of a
given node d = 3 is 24 and by neighbors of node i we considered its nearest and next-nearest neighbors. Despite the
fact that the dynamics of the model is based on complementary interactions and involve mirror images, the neighbors
of a given site in a subnetwork correspond to the populations in the same state belonging to its neighborhood, as
defined above. The average values used for the majority of graphs as well as the distributions, over 1000 samples were
obtained, discarding the 1000 initial time steps, in order to guarantee that the samples correspond to thermalized
states on the different regimes.
1
0,8
0,6
0,4
0,2
/
Ν
Β
#
0
0
0,1
0,2
x
0,3
0,4
FIG. 1. The average population densities as a function of the parameter control of initial configurations (x). The symbols
correspond to: ((cid:13)) B0; ((cid:3)) B1 and (△) B2 concentrations. There is a remarkable change in the behavior of the densities for
x ∼ 30 separating the stable and chaotic phases.
The transition region between stable to chaotic behaviors for this set of parameters is located in the vicinities of
x ∼ 30: the stable phase corresponded to the region of x < 30, and the chaotic associated to the region of x > 30.
Figure 1 shows that the stable phase is dominated by populations in low concentration (B = 0), while in the chaotic
phase the density of the three types of populations are of the same order of magnitude and therefore, 2/3 of the nodes
are activated (B = 1 and B = 2) corresponding to a pathological state as previously suggested [9]. All values shown
in Figure 1 correspond to the values of the densities calculated after thermalization.
The degree distributions of the three subnetworks (B0, B1 and B2) in the stable phase are finite and exponential
distributions, as shown in Figure 2(a). The positive exponent for B0 distribution indicates a high probability that any
given node belonging to this subnetwork will be linked to almost all its neighbors, thus generating large domains of
low-activity nodes on the 3-D lattice. The negative slope obtained for B1 and B2 distributions indicate that in both
1
0,01
)
K
(
P
0,0001
1e-06
0,2
0,15
)
K
(
P
0,1
0,05
6
1e-08
0
2
4
6
8 10 12 14 16 18 20 22 24
K
0
0
2
4
6
8 10 12 14 16 18 20 22 24
K
(a)
(b)
FIG. 2. The degree distribution P (K) × k of the three subnetworks in the stable x = 0.20 and chaotic x = 0.40 phases. (a)
Stable phase: all distributions are exponential and the slopes obtained are a= 0.83, -1.03 and -0.95 for subnetworks B0, B1
and B2 respectively. (b) Chaotic phase: all distributions are Poissonian. The different symbols refer to: ((cid:13)) B = 0; ((cid:3)) B = 1
and (△) B = 2.
subnetworks the active nodes have a high probability of being linked to very few neighbors. The stable regime would
correspond to atypical states of the immune system with very few active populations, a feature that makes it difficult
to maintain a memory regarding the perturbation history of the system, and thus explains why the mice experiments
[11] could not be reproduced in this region of the parameter space.
As shown in Figure 2(b), in the chaotic regime the degree distributions are similar, finite and Poissonian (or bell-
shaped) with an average number of connections hKi equal to 8, 7 and 10 for B0, B1 and B2 subnetworks, respectively.
Therefore, theses subnetworks are random networks [20] as will be the entire network; this regime, as mentioned above
[9] would also correspond to a pathological state of the immune system, in which the order of 2/3 of the populations
are randomly activated corresponding to an overactive state equivalent to that of septicemia.
1
0,1
)
K
(
P
0,01
0,001
0,0001
0,2
0,15
)
K
(
P
0,1
0,05
1e-05
0
2
4
6
8 10 12 14 16 18 20 22 24
K
0
0
2
4
6
8 10 12 14 16 18 20 22 24
K
(a)
(b)
FIG. 3. Degree distribution P (K) × K at x = 0.30. The degree distribution of subnetwork B0 (a) presents mixed behavior:
half exponential and half Poissonian. (b) B1 and B2 distributions are pure Poissonian distributions. The different symbols
refer to: ((cid:13)) B = 0; ((cid:3)) B = 1 and (△) B = 2.
In the transition region (x = 0.30) we find a completely different behavior of the B0 distribution with respect
to B1 and B2 distributions, as shown in Figure 3. The B0 degree distribution indicates mixed behavior, exhibiting
characteristics of both regimes, stable and chaotic. For K < 14 the B0 distribution is Poissonian and for K ≥ 14 it
is exponential; thus, in the transition region the B0 nodes have a large probability of being linked to more than 50%
of its neighbors, forming large clusters. However, embedded in this structure there is a random network of B0 nodes
with low connectivity (K < 14) exhibiting a characteristic connectivity or an average value of K. As far as we know,
this is the first time that such a type of mixed behavior of the degree distribution of a network has been reported on
the literature. The degree distributions for B1 and B2 shown in Figure 3(b) although showing log-normal behavior
can be identified as Poisson-like distributions with long tails, corresponding to random subnetworks with few highly
connected nodes (long tails). These random subnetworks have active nodes with low connectivity, and therefore, under
perturbation would be able to change more easily than highly connected nodes to incorporate information into the
subnetwork. The average degree for these distributions (5 and 6 for B1 and B2 distributions respectively) is smaller
than the average K obtained in the distributions of the chaotic regime, which in the case of B2 is almost double.
7
As reported in the literature, many biological networks are scale-free and this structure would guarantee the identity
and robustness of the network [14, 15, 21]. The same behavior was obtained by Burns and Ruskin [22] for the degree
distribution of a different immune network model also based on shape-space formalism that was proposed to describe
the development of the repertoire of immune cells. The BSP model is a different model describing only B-cell
population interactions based on non-local rules inspired by the functioning of the immune system. The networks
it generates under certain conditions are able to reproduce the dynamical memory and behavior observed in real
immune systems. Poissonian behavior predominates in the degree distribution for the subnetworks in transition and
the chaotic phases. These results are very interesting since they suggest that random structures are always present
in the active subnetworks of these regimes and would be responsible for the dynamic allocation of memory observed
in the transition region. The randomly attached active sites of low connectivity allow for rapid changes to add or
discard information. The plasticity, which permits the memory of previous antigen (perturbation) history to be
maintained, only observed on the transition region, will also depend on the behavior of the degree distribution of
low concentration subnetworks exhibiting a mixed behavior (exponential and pure Poissonian) 3. In order to achieve
the plasticity necessary to maintain the history of the perturbation (antigens) memory, it is necessary for the low
connectivity subnetwork to have a random structure embedded on one that is very connected and almost regular:
low connectivity sites would allow for rapid changes following the active populations while the high connected ones
would keep the robustness and the integrity of the system.
In the chaotic regime the fact that all distributions
are Poissonian with a relatively low average connectivity allows for rapid changes in all subnetworks, and since the
densities of different populations are of the same order of magnitude rapid changes would very easily lead to the
over-activation of the populations. Therefore, the randomness on the degree distribution is an important feature for
the immune network [2] to maintain a memory of the previous history of antigen presentations. The exponential
behavior observed on the degree distribution for all subnetworks in the stable regime do not allow the memory to
be maintained for two reasons: overall, the high density of low concentration sites is highly connected, which is an
unfavorable aspect regarding the necessary changes to incorporate information, and the very low density of active
sites does not allow for the accumulation of information.
25
20
)
K
(
15
n
n
K
10
5
0
0
(a)
25
20
)
K
(
15
n
n
K
10
5
0
0
(b)
5
10
K
15
20
25
5
10
K
15
20
25
)
K
(
n
n
K
24
20
16
12
8
4
0
0
(c)
2
4
6
8 10 12 14 16 18 20 22 24
K
FIG. 4. Average degree of the nearest neighbors Knn(K) for nodes with degree K. An assortative mixed behavior is observed
for sub networks ((cid:13)) B0; ((cid:3)) B1 and (△ B2) at: (a) Stable phase (x = 0.20), (b) Transition region (x = 0.30) and (c) the
chaotic region (x = 0.40), for which the slope of the solid guide line is 0.44.
We have also studied the behavior of the average degree Knn(K) of the neighbors of the nodes with degree K.
Overall, the results indicate that there is a positive correlation between the neighbors or assortative mixing properties,
i.e., nodes with the same type of connectivity (low and high) tend to cluster together. Moreover, results indicate that
there is an overall linear growth of Knn(K) as a function of K, but behavior is different for all three phases. In the
8
stable regime Knn(K) grows almost linearly (positive correlation) for K ≥ 14, K ≤ 3 and K ≤ 7 for B0, B1 and B − 2
distributions, respectively, but decreasing very rapidly (negative correlation) afterwards in the last two distributions
(Fig. 4(a)). In other words the positive correlation grows up to a critical value (Kc) of connectivity and becomes
negative for K ≥ Kc. At the transition we observed that the average degree for B0 distribution behaves linearly for
all K, although with different slopes in at least two regions; the behavior of B1 and B2 persists as in the ordered
phase, but in this case they grow linearly for K ≤ 12 and K ≤ 14 respectively, with a subsequent decrease, as in
the previous regime (Fig. 4(b)). The existence of cut-off degrees is characteristic of finite random networks, however
the linear growth has not previously been observed. In the chaotic region (x = 0.40) Knn(K) grows linearly with
K ((see Figure 4(c)) for all subnetworks, as expected, since in this region, the degree distribution is Poissonian and
the densities of nodes for all subnetworks are of the same magnitude. The positive correlation would allow for the
over-activation of nodes as observed.
For the same set of parameters N = 283, in the transition and chaotic regions (x = 0.3 and x = 0.4 respectively) we
have estimated the mean shortest path L for the three subnetworks. Due to the computational cost it is difficult to
perform this calculation with the necessary statistics, which is why here we present the results obtained for a simple
sampling. At the transition region we have obtained L = 10.72 (N = 14433) for B0, L = 13.05 (N = 2872) for B1
and L = 11.60 (N = 4338) for B2 and for the chaotic regime (x = 0.40) the results are: L = 11.38 (N = 7471) for
B0, L = 11.76 (N = 6289) for B1 and L = 11.19 (N = 8185) for B2. In all cases we observe that L ≪ N (number
of nodes in the subnetwork) but overall logN ≤ L ≤ L1/d, i.e. the behavior between regular lattice and random
networks. Therefore, L ≪ N is not necessarily a signature of small-world properties as suggested by other biological
networks [23], but in our case it is a signature of the existence of very connected nodes. We should expect small-word
properties as in other random networks, however due to the computational cost we have not looked for evidence of
such properties.
When we compute the clustering coefficient C as a function of x for the three subnetworks (shown in Figure 5),
we observe that for the subnetwork B0 C is approximately constant, with a slight increase from 0.38 to 0.40 in the
transition region. This result indicates a strong aggregation between the populations at a low concentration, i.e., on
average around 40% percent of the neighbors are neighbors among themselves. However, the behavior of the clustering
coefficients for subnetworks B1 and B2 differ completely from that of B0: for small x in both cases the clustering
coefficient is close to zero and start to increase around x = 0.20, reaching its maximum value (∼ 0.4) for x > 0.3.
The increase of the clustering coefficient implies an increase in the aggregation of these subnetworks, reflecting an
increase in the concentrations of active sites with intermediate and high concentrations as x increases (Figure 1). We
have also observed that the cohesiveness between different populations is always less significant than the cohesiveness
among the elements of the same population (results not shown).
0,5
0,4
0,3
0,2
0,1
>
)
x
(
C
<
0
0
0,1
0,2
x
0,3
0,4
FIG. 5. The global clustering coefficient C as a function of x for the subnetworks: ((cid:13)) B0; ((cid:3))B1 and (△)B2.
While the average clustering coefficient is almost constant as a function of K in the chaotic and transition regions, it
exhibits a power-law behavior in the stable region (x = 0.2), as shown in Figure 6(a). This behavior indicates that the
subnetworks in the stable regime have hierarchical structures [18] with exponents that are smaller than unity (0.29 for
B1 and 0.40 for B2) as observed in the case of the internet [24]. Hierarchical structures go beyond simple clustering by
including organization at all scales in the network simultaneously. Trees, in which the nodes are connected, generally
possess common ancestors that are topologically closer to them than nodes, which are not closely related. Since the
numbers of populations with intermediate or high concentration are small in the stable regime, this would explain
why under perturbation the system changes, incorporating information regarding perturbation and losing information
on the previous perturbation, do not maintain any memory regarding its history [12, 13].
Figure 6(b) shows the constant behavior of the average clustering coefficient as a function of K in the chaotic region.
Constant behavior is characteristic of random networks where the clustering coefficient is Ci = p where p = hki/N is
the probability of connection between two nodes. For random networks the small aggregation is due to small p [20].
Previous results indicate that in the chaotic region the subnetworks are random networks.
9
V. CONCLUSIONS
In this work, from the viewpoint of a complex network we have analyzed an immune network model, which was able
to reproduce the behavior of the immune systems of mice. Indeed, the immune network is complex, and contrary to
what has been reported in the literature regarding other biological networks, this biological network is characterized
by random behavior rather than scale-free [14].
1
>
)
K
C
<
(
1
(a)
1
0,8
0,6
0,4
0,2
>
)
K
C
<
(
10
K
100
0
0
2
4
6
8 10 12 14 16 18 20 22 24
K
(b)
FIG. 6. Average clustering coefficient hC(K)i as a function of the connectivity K for (a) Stable (x = 0.20) and (b) Chaotic
(x = 0.40) phases, where ((cid:13)) corresponds to B0, ((cid:3)) to B1 and (△) B2 sub networks, respectively.
In fact, the networks obtained in the three dynamical regimes observed in the model are composed of three different
subnetworks of low, intermediate and high concentrations. In this work we have investigated the behavior of the three
subnetworks in the different regimes. We observed the existence of random characteristics in the three regions:
in
the ordered phase the degree distributions are purely exponential; in the transition region while B0 mix exponential
(K ≤ 14) and Poisson (K > 14), the B1 and B2 are Possonian; and in the chaotic region the distributions are purely
Poissonian. In reality, the immune network seems regulated by random subnetworks of active populations (B1 and
B2). The randomness of these structures formed by nodes with low average connectivity (with respect to the maximum
number of neighbors) allow for rapid changes that add or discard network information. However the necessary plasticity
that would allow for adaptation and maintainance of the networks identity and memory regarding its history in terms
of perturbations (or antigen presentations) depends on the structure of the low concentration population subnetwork.
The structure that allows the reproduction of immune system behavior in mice is composed of random substructures
embedded in one that is very connected:
low concentration sites with random low connectivity will allow for the
incorporation of new information (new active populations in the network) leading to a different configuration that
when compared to the previous one has added relevant new information and discarded the unnecessary.
Another important result that differs from those reported in the literature, is the fact that Knn(K) grows linearly
as a function of K. For stable and transition regions the active population subnetworks show a linear growth (positive
correlation) with a cut-off Kc, as expected for finite random networks, followed by a negative correlation. However,
B0 distributions do not have a cut-off; in the chaotic region, all distributions exhibit a linear growth of Knn(K) as
a function of K. Put another way, while the B0 subnetworks exhibit assortative mixing in all phases, there is an
interplay between assortative and disassortative mixing in B1 and B2 subnetworks at stable and transition regions,
depending on the value of K. If K ≤ Kc there is a positive, linear degree-correlation, but for K > Kc the correlation
is negative. There is a maximum threshold value in the transition region but that disappears in the chaotic region
giving way to pure assortative behavior. This suggests that although random, the networks obtained with this model
are much more complex than those thus far reported in the literature. We believe that the differences between our
results and those from the literature arise from the fact that in this study the networks topology reflects the dynamics
inspired in the biological process attributed to the immune system, while the majority of biological networks that
have been studied neither include nor reflect the dynamics of the process, since in many cases it is unknown.
ACKNOWLEDGMENTS
10
We thank Silvio Ferreira Jr.
for his helpful discussions and suggestions. This study received financial support
from the following Brazilian institutions: the Conselho Nacional de Desenvolvimento Cient´ıfico e Tecnol´ogico (CNPq,
http://www.cnpq.br), the Coordena¸cao de Aperfeioamento de Pessoal de Nivel Superior (CAPES, http://www.capes.gov.br)
and the Funda¸cao de Amparo Ciencia e Tecnologia do Estado de Pernambuco (FACEPE, http://www.facepe.br-
Pronex-EDT 0012-05.03/04 and Pronex-APQ 0203-1.05/08).
[1] C. A. Janeway, P. A. Traver, M. Walport, and J. Capra, Immunobiology: The Immune System In Health And Disease
(Garland Science Publishing, NY, 2009)
[2] N. K. Jerne, Ann. Immuno. (Inst. Pasteur) 125 C, 372 (1974)
[3] A. Coutinho, Immunol. Rev. 110, 63 (1989)
[4] D. Holmberg, Anderson, L. Carlsson, and S. Forsgren, Immunol. Rev. 110, 89 (1989)
[5] I. Lundkvist, A. Coutinho, F. Varela, and D. Holmberg, Proc. Natl. Acad. Sci. USA 86, 5074 (1989)
[6] D. Stauffer and G. Weisbuch, Physica A 180, 42 (1992)
[7] R. J. D. Boer, L. A. Segel, and A. S. Perelson, J. Theor. Biol. 155, 295 (1992)
[8] R. M. Z. dos Santos and A. T. Bernades, Physica A 219, 1 (1995)
[9] R. M. Z. dos Santos, Physica A 196, 12 (1993)
[10] A. T. Bernardes and R. M. Z. dos Santos, J. Theor. Biol. 186, 173 (1997)
[11] R. M. Z. dos Santos and A. T. Bernades, Phys. Rev. Let. 81, 3034 (1998)
[12] R. M. Z. dos Santos and M. Copelli, Brazilian Journal of Physics 33, 628 (2003)
[13] M. Copelli, R. M. Z. dos Santos, and D. A. Stariolo, Eur. Phys. J. B 34, 119 (2003)
[14] R. Albert and A.-L. Barab´asi, Rev. Mod. Phys. 74, 47 (2002)
[15] H. Jeong, S. P. Mason, Z. N. Oltvai, and A. L. Barab´asi, Nature 411, 41 (2001)
[16] A. S. Perelson and G. F. Oster, J. Theor. Biol. 81, 645 (1979)
[17] A. T. Bernardes and R. M. Z. dos Santos, Int. J. Mod. Phys. C 12, 1 (2001)
[18] E. Ravasz and A.-L. Barab´asi, Phys. Rev. E 67 (2003)
[19] A. Barrat, M. Barth´elemy, and A. Vespignani, Dynamical Processes on Complex Networks (Cambridge Univ. Press, 2009)
[20] P. Erdos and A. R´enyi, Publ. Math. 6 (1959)
[21] B. A-L, Z. N. Oltvai, and S. Wuchty, Lect. Notes. Phys. 650, 443 (2004)
[22] H. J. Ruskin and J. Burns, Physica A 365, 549 (2006)
[23] D. J. Watts and S. H. Strogatz, Nature 393, 440 (1998)
[24] A. V´azquez, R. Pastor-Satorras, and A. Vespignani, Phys. Rev. E 65, 066130 (2002)
|
1009.3085 | 1 | 1009 | 2010-09-16T03:47:13 | Approximation scheme based on effective interactions for stochastic gene regulation | [
"physics.bio-ph",
"q-bio.MN"
] | Since gene regulatory systems contain sometimes only a small number of molecules, these systems are not described well by macroscopic rate equations; a master equation approach is needed for such cases. We develop an approximation scheme for dealing with the stochasticity of the gene regulatory systems. Using an effective interaction concept, original master equations can be reduced to simpler master equations, which can be solved analytically. We apply the approximation scheme to self-regulating systems with monomer or dimer interactions, and a two-gene system with an exclusive switch. The approximation scheme can recover bistability of the exclusive switch adequately. | physics.bio-ph | physics |
Approximation scheme based on effective interactions for stochastic gene regulation
Jun Ohkubo∗
Graduate School of Informatics, Kyoto University,
36-1, Yoshida Hon-machi, Sakyo-ku, Kyoto-shi, Kyoto 606-8501, Japan
(Dated: July 23, 2018)
Since gene regulatory systems contain sometimes only a small number of molecules, these systems
are not described well by macroscopic rate equations; a master equation approach is needed for such
cases. We develop an approximation scheme for dealing with the stochasticity of the gene regulatory
systems. Using an effective interaction concept, original master equations can be reduced to simpler
master equations, which can be solved analytically. We apply the approximation scheme to self-
regulating systems with monomer or dimer interactions, and a two-gene system with an exclusive
switch. The approximation scheme can recover bistability of the exclusive switch adequately.
I.
INTRODUCTION
Recently, stochastic nature in small systems has at-
tracted many attentions [1 -- 3]. One of the interesting
examples of the stochasticity is a gene regulatory sys-
tem;
it has been known experimentally that the gene
regulatory systems show various phenomena caused by
intrinsic noise [4, 5]. The gene regulatory systems basi-
cally consists of genes, RNAs, and proteins. The genes
could sometimes be activated or repressed by regulatory
proteins known as transcription factors. The number of
regulatory proteins is sometimes very small, and there
are large fluctuations. From a theoretical point of view,
the gene regulatory systems have been studied a lot us-
ing Monte Carlo simulations (e.g., [6, 7]). In addition,
in order to gain insights into mechanisms or functions
of the gene regulatory systems, many analytical studies
have been done [8 -- 18]. For example, if we consider a
self-regulating gene with monomer binding interactions,
an exact solution has been already known [12]. When
one considers more complicated systems, some approx-
imations are needed. Such approximations have also
been developed; Fokker-Planck or Langevin equation ap-
proach [8 -- 10], a variational approach [11, 14, 15], and
self-consistent proteomic field approximation [13].
A gene regulatory system with only two genes and feed-
back mechanisms has been studied a lot because it plays
an important role as a genetic switch; two distinct sta-
ble states emerges, and they could be switched either
spontaneously or by external signals.
In mathematical
description for the gene regulatory systems, the RNAs
are sometimes neglected for simplicity, and only genes
and regulatory proteins are considered. When we con-
struct a macroscopic rate equation, in which fluctuations
in protein copy numbers or gene expression states are
neglected, the analysis for the rate equation tells us the
following facts: A system with two mutually repressing
genes shows a bistability, and cooperative binding of reg-
ulatory proteins is important for making the bistability
[19, 20]. Here, the cooperative binding means that com-
∗Email address: [email protected]
binations of two or more proteins need to activate or
repress genes. The macroscopic rate equation gives mul-
tiple stable solutions, and each solution corresponds to a
stable state of the gene regulatory systems, which causes
the bistability. Hence, for the cooperative binding cases,
it may be enough to use the macroscopic treatments in
order to investigate qualitative behavior of the bistabil-
ity. However, other studies have shown that a so-called
exclusive switch shows a bistability even when the macro-
scopic rate equations have only one solution [21 -- 23]. Al-
though the bistability has been confirmed numerically us-
ing Monte Carlo simulations, no exact or approximated
analytical treatment has been proposed yet, to the best
of our knowledge.
In the present paper, we develop a new approximation
scheme for gene regulatory systems. In the approxima-
tion scheme, there is no need to use continuous descrip-
tion such as Fokker-Planck or Langevin equations, and
hence the smallness or discrete properties of the system
are not neglected. The basic idea of the approximation
is similar to the "self-consistent proteomic field approx-
imation" developed by Walczak et al.
[13]. In the self-
consistent proteomic field approximation, a joint proba-
bility for all genes is approximated as a product of prob-
ability distributions for each gene, and then the interac-
tions between genes and regulatory proteins can be eval-
uated 'exactly' in this approximation. In [13], only toggle
switches, which consist of two genes, have been studied;
as denoted in the discussions in [13], further approxima-
tion would be needed for self-regulating systems. We here
extend the concept of [13], and develop a more applicable
approximation scheme; the interactions between genes
and regulatory proteins are approximated firstly, and an
effective interaction is introduced. The new approxima-
tion scheme would be useful to treat more complicated
cases, such as the exclusive switch. The new approxi-
mation scheme enables us to give analytical expressions
for probability distributions of the numbers of proteins,
without loss of the discreteness property of the system.
The effective interactions are estimated self-consistently.
We will demonstrate the usefulness of the approximation
scheme by using self-regulating systems and the exclusive
switch without cooperative interactions.
The present paper is constructed as follows. In Sec. II,
we give a brief review of a stochastic model for gene reg-
ulation. In Sec. III, self-regulating systems are studied.
Section III C gives one of the important results in the
present paper, in which our approximation scheme is pro-
posed. The proposed approximation scheme is applied to
the exclusive switch in Sec. VI. Section V is concluding
remarks.
II. STOCHASTIC MODEL FOR GENE
REGULATION
We here briefly review the basic biology of genetic regu-
latory system and a simplified stochastic model, for read-
ers' convenience.
A gene regulatory system consists of many compo-
nents, such as genes, RNAs, and proteins. The tran-
scription of a gene is initiated by a binding of RNA poly-
merase to a promoter site of the gene in the DNA. The
binding of regulatory proteins (or molecules), so-called
transcription factors, can sometimes regulate the tran-
scription initiation. These regulatory proteins bind to
own target operator sites, and they sometimes act as re-
pressors (which repress the transcription) or activators
(which enhance the transcription) of the transcription.
When the RNA polymerase binds to a gene, the gene
sequence is copied into a messenger RNA (mRNA), and
the mRNA is translated into a protein molecules by a
ribosome enzyme complex. The produced proteins are
important to determine the phenotypic behavior of the
cell. In addition, regulation of transcription is one im-
portant way of controlling the phenotypic behavior, and
sometimes the produced proteins can become regulatory
signals for genes.
Although all of the above reactions would be important
for the gene regulatory systems, the mRNA is sometimes
neglected in stochastic modelling for simplicity. That is,
the translation from mRNAs to proteins are straightfor-
ward, and then we assume that an activated gene directly
increases the number of proteins. In addition, we con-
sider that a repressed gene cannot produce any proteins,
which makes analytical treatments much simpler [12].
In the present paper, all regulatory proteins act as re-
pressors. If regulatory proteins are not binding to a gene,
then we call a state of the gene as 'ON' state; if not, the
gene is in 'OFF' state. A gene in OFF state cannot pro-
duce any proteins, as we assumed above.
III. SELF-REGULATING SYSTEM
In this section, we will explain a new approximation
scheme using a simplest model, i.e., a self-regulating sys-
tem. Exact solutions for the self-regulating system with
monomer interactions have already been known. After
reviewing the exact solutions, we will propose a new ap-
proximation scheme. The new approximation scheme will
2
FIG. 1: A schematic illustration of the self-regulating gene.
be applied to the self-regulating systems with monomer
and dimer interactions, respectively.
A. Model
At first, we give a brief explanation for a self-regulating
system. In the self-regulating system, there is only one
gene, and it produces proteins. The produced proteins
are considered as regulatory proteins for the gene, and
the regulation is a repressed one.
In this sense, there
is a self-regulation mechanism. Figure 1 shows the self-
regulating system. When the gene is in ON state, it
produces proteins with rate g. The degradation rate of
the regulatory proteins is k. The regulatory proteins can
bind the gene with rate function H(n), where n is the
number of 'free' regulatory proteins. The function H(n)
can be a complicated function of the regulatory proteins;
e.g., H(n) = hn for monomer interactions, and H(n) =
hn(n − 1)/2 for dimer interactions, where h is a rate
for the binding. f is the rate with which the regulatory
protein is released from the repressor site of the gene.
B. Exact solution for monomer interactions
We here consider a simplest interaction case, i.e., a
monomer interaction case. Hence, H(n) in Fig. 1 is writ-
ten as hn, as discussed in Sec. III A. For the monomer in-
teraction cases, exact solutions have already been known
[6, 12]. In order to compare our approximation scheme,
which will be proposed in Sec. III C, we here briefly re-
view the exact solutions.
In order to make analytical treatments simpler, one
assumption should be included [6]; i.e., one of the pro-
teins in ON state is inert, and then the protein cannot
be degraded or repress to the gene. Hence, there are
a little difference between usual stochastic simulations
and this analytical treatment. However, it has already
been discussed that this assumption alter only for lower
numbers of proteins, and actually it gives quantitatively
good results [6]. Hence, we here employ this assump-
tion. Figure 2 shows the transition scheme for the usual
stochastic simulations and the analytical treatment. αn
and βn correspond to probabilities with which there are n
example, the number of 'free' regulatory proteins is
α′
n =
∂ n
∂z n α′(z)(cid:12)(cid:12)(cid:12)(cid:12)z=0
.
3
(5)
The probability with which the gene is in ON state is
given as α′(1); the number of total proteins in the system
is given as
hni =
∂α′(z)
∂z
(cid:12)(cid:12)(cid:12)(cid:12)z=1
+ 1 × α′(1) +
,
(6)
∂β(z)
∂z (cid:12)(cid:12)(cid:12)(cid:12)z=1
where the second term in r.h.s means a contribution from
the inert protein in ON state.
Eqs. (1) and (2) can be rewritten as two differential
equations in terms of the generating functions;
∂α(z)
∂t
= (z − 1)(cid:20)gα′(z) − k
∂α′
∂z (cid:21) − hz
∂α′
∂z
+ f β(z),
∂t
∂β(z)
= −(z − 1)k
∂β
∂z
After some calculations,
generating functions are obtained as follows [6, 12]:
− f β(z).
∂α′
∂z
'stationary' solutions for the
+ hz
(7)
(8)
α′ ex(z) =Aex F (1 + aex, 1 + bex, N ex(z − zex
∂α′ ex(z)
0 )),
(z − 1)(cid:20)gα′ ex(z) − k
∂z
βex(z) = −
1
f
h
f
+
z
∂α′ ex(z)
∂z
,
(cid:21)
(9)
(10)
where the super-script 'ex' means 'exact solutions', and
, N ex =
g
k + h
,
zex
0 =
aex =
k
k + h
f
k
,
bex =
f
k + h
+ (1 − zex
0 )N ex.
(11)
Aex is the normalization constant, which is determined as
α′ ex(1) + βex(1) = 1. F (p, q, r) is the Kummer confluent
hypergeometric function,
F (p, q, r) ≡
∞
Xn=0
(p)n
(q)n
rn
n!
,
(12)
where (p)n = p(p + 1)(p + 2) · · · (p + n − 1).
Details of the characteristics of the exact solution are
written in [6]. For example, the probability distributions
for the numbers of 'free' regulatory proteins are
α′ ex
n =
(N ex)n (1 + aex)n
(1 + bex)n
Aex
n!
× F (1 + aex + n, 1 + bex + n, −N exzex
Aex
(N ex)n
0 ),
f (cid:20)((1 − zex
0 )N ex − bex)
n!
(13)
×
(1 + aex)n
(1 + bex)n
+bex (N ex)n
n!
F (1 + aex + n, 1 + bex + n, −N exzex
0 )
(aex)n
(bex)n
F (aex + n, bex + n, −N exzex
0 )(cid:21) ,
(14)
FIG. 2: Transition scheme for simulations and analytical
calculations for monomer interaction cases. αn and βn cor-
respond to probabilities with which there are n regulatory
proteins for ON and OFF states, respectively. In the analyt-
ical calculations, one of the proteins is considered as an inert
one when the gene is in ON state, and the inert protein is
also included in αn; the number of 'free' regulatory proteins
in ON state is n − 1.
regulatory proteins for ON and OFF states, respectively.
In the usual stochastic simulations, the degradation rate
of the proteins, i.e., the change from αn to αn−1, is pro-
portional to the number of proteins, n. In contrast, the
above assumption means that the rate from αn to αn−1
in the analytical treatment is proportional to n − 1, not
to n.
In this assumption, the master equations are given as
n
dα′
dt
dβn
dt
=g[α′
n−1 − α′
n] + k[(n + 1)α′
n+1 − nα′
n]
− hnα′
n + f βn,
= + k[(n + 1)βn+1 − nβn]
+ hnα′
n − f βn,
(1)
(2)
where α′
n is a probability with which there is n 'free'
regulatory proteins for ON state; αn+1 ≡ α′
n. Note that
the inert protein is not a 'free' regulatory protein, and
α′
n does not include the inert protein.
In order to solve eqs. (1) and (2), it is useful to define
the following generating functions;
α′(z) =
β(z) =
∞
Xn=0
Xn=0
∞
α′
nz n,
βnz n.
(3)
β ex
n =
(4)
Using the generating functions, various information
about the self-regulating system can be obtained. For
where f = f /(k + h).
According to the following discussions, we here set the
effective interaction h as
C. Approximation scheme
h = hhniα′ ,
(20)
4
In Sec. III B, we analyzed the self-regulating system
with monomer interactions. In the system, the interac-
tion factor, H(n), is given as hn, and actually this sim-
ple form of the interaction enables us to obtain the exact
solutions. If we consider different types of interactions,
such as dimer ones, exact solutions have not been known
yet.
Here, we propose a new approximation scheme. The
key of the approximation scheme is to use "an effective
interaction". Although the new approximation scheme is
similar to the self-consistent proteomic field approxima-
tion in [13], the new one is more applicable; it is applica-
ble even to the self-regulating systems or exclusive switch
cases, as shown later. The effective interaction means
that the interaction factor in Fig. 1, H(n), is replaced
as a scalar value; i.e., H(n) = h. Here, the effective in-
teraction h is not a function of the regulatory proteins.
Hence, the master equation for this approximated system
is written as follows:
n
dα′
dt
dβn
dt
=g[α′
n−1 − α′
n] + k[(n + 1)α′
n+1 − nα′
n]
− hα′
n + f βn,
= + k[(n + 1)βn+1 − nβn]
+ hα′
n − f βn.
(15)
(16)
Because the interaction factor H(n) = h has a simple
form, the analytic solution can be easily calculated by
using the generating function approach. Putting the left-
hand sides of Eqs. (15) and (16) as zero and rewriting
Eqs. (15) and (16) in terms of the generating functions
α′(z) and β(z), stationary solutions for the generating
functions are obtained as follows:
α′(z) =A F (a, b, N (z − 1)),
(17)
β(z) =(cid:18)1 +
h
f(cid:19) A F (a − 1, b − 1, N (z − 1)) − α(z),
(18)
where A = f /(f + h) and
N =
g
k
,
a = 1 +
f
k
,
b = 1 +
f + h
k
.
(19)
A remaining task is to determine the effective inter-
action h. For the self-regulating system with monomer
interactions, the binding of the regulatory proteins oc-
curs only when the system is in ON state. Hence, the
number of proteins, which can be attached to the bind-
ing site, should be equal to the number of free proteins
for ON state.
where hniα′ is the expectation of the number of free reg-
ulatory proteins under a condition that the gene is in ON
state (conditional expectation). Because it is possible to
evaluate the conditional expectation using the generating
function (Eq. (17)) as
hniα′ ≡
1
α′(1)
∂
∂z
=
g(k + f )
k(f + f + h)
,
(21)
α′(z)(cid:12)(cid:12)(cid:12)(cid:12)z=1
we obtain the following self-consistent equation by insert-
ing Eq. (20);
hhniα′ = h
g(k + f )
k(f + f + hhniα′ )
.
(22)
Solving Eq. (22), we finally obtain
hhniα′ =
−(k2 + kf ) +p(k2 + kf )2 + 3khg(k + f )
.
2k
(23)
Once the effective interaction h is determined, all sta-
tistical properties related to the number of regulatory
proteins are evaluated from the generating functions
(Eqs. (17) and (18)). For example, the probability dis-
tributions for the numbers of free proteins are
F (a + n, b + n, −N ),
(24)
n =AN n (a)n
α′
(b)n
βn =(cid:18)1 +
h
f(cid:19) AN n (a − 1)n
(b − 1)n
× F (a − 1 + n, b − 1 + n, −N ) − α′
n.
(25)
D. Results for monomer interactions
For monomer interaction cases, the exact solutions are
obtained. Hence, we here compare the exact results
and approximate results obtained by the approximation
scheme.
For the comparison, we here introduce rescaled param-
eters as follows [6]:
ω =
f
k
, Xeq =
f
h
, Xad =
g
2k
,
(26)
and for simplicity, we set k = 1 in all numerical evalu-
ations. These rescaled parameters are helpful to under-
stand properties of the genetic switch. The parameter
Xad characterizes the synthesis/degradation processes,
and large Xad would give a large average number of pro-
teins. Xeq is related to the equilibrium constant of the
binding/unbinding process. Finally, ω is a parameter
called "adiabaticity parameter". ω measures how rapidly
Case A: Exact
Case B: Exact
Case A: Approx.
Case B: Approx.
0 10 20 30 40 50 60 70 80 90 100
n
Exact
Approx.
(a)
y
t
i
l
i
b
a
b
o
r
p
(b)
0.10
0.08
0.06
0.04
0.02
0.00
13
12
11
10
9
8
>
n
<
7
0.0
2.0
4.0
ω
6.0
8.0
10.0
FIG. 3: Comparison between results from the exact solu-
tions and the approximation scheme. (a) The probability dis-
tributions of the number of proteins. Case A: Xeq = 10.0,
Xad = 10.0, ω = 0.01. Case B: Xeq = 50.0, Xad = 50.0,
ω = 10.0. For Case A, it is difficult to see the difference be-
tween the exact solution and approximate solution. (b) The
average number of proteins. Xeq = 10.0 and Xad = 10.0, and
only the value of ω was changed.
the gene can equilibrate in a gene state. If ω is small, the
synthesis/degradation behaves almost like an indepen-
dent birth and death process, and there would be two
peaks corresponding to the binding/unbinding states, re-
spectively. For details of these parameters, e.g., see [6].
Firstly, the probability distributions of the number
of protein were compared. Figure 3(a) shows the re-
sults. Here, we performed two cases: In Case A, we set
Xeq = 10.0, Xad = 10.0, ω = 0.01; in Case B, Xeq = 50.0,
Xad = 50.0, ω = 10.0. For Case A, the exact solution
and the approximate one give a good agreement, and
it is difficult to see the difference. Although there are
quantitative differences between the exact and approx-
imate solutions for Case B, the approximation scheme
gives qualitatively good result despite of the rough ap-
proximation. Figure 3(b) shows the average number of
proteins (Eq. (6)) for various values of ω when Xeq = 10.0
and Xad = 10.0. If ω is small, the approximation scheme
gives quantitatively good results. Even in the large ω
5
FIG. 4: Transition scheme for simulations and analytical
calculations for dimer interaction cases.
case, the difference between the exact and approximate
results are less than 1.
E. Results for dimer interactions
As a second example, a self-regulating system with
dimer interactions are considered. In this case, the tran-
sition scheme for analytical calculations are different from
the monomer interaction cases; see Fig. 4. In this case,
the master equations are the same as Eqs. (15) and (16),
but the effective interaction should be set as
h = h
hn(n − 1)iα′
2
,
(27)
and we should interpret α′
n as αn+2 = α′
n.
Using the similar procedure written in Sec. III C, the
effective interaction h is obtained by solving the following
self-consistent equation:
h
hn(n − 1)iα′
2
=
h
2
1
α′(1)
∂2
∂z2 α′(z)(cid:12)(cid:12)(cid:12)(cid:12)z=0
.
(28)
Since it is a little complicated task to obtain the analyt-
ical expression for the effective interaction h, we numer-
ically solved the self-consistent equation (Eq. (28)).
For the dimer interaction cases, we have not obtained
any exact solution. Hence, results of the approximation
scheme were compared with those of the Monte Carlo
simulations.
Figure 5 shows the probability distributions. In Case
A, we used the following rescaled parameters: Xeq =
10.0, Xad = 10.0, ω = 0.01; in Case B, Xeq = 50.0,
Xad = 50.0, ω = 10.0. The numbers of the Monte Carlo
steps are over 107 for Case A, and 108 for Case B. Obvi-
ously, the approximation scheme gives qualitatively good
results; although the shapes of the distributions and the
positions of peaks are slightly different, the number of
peaks are the same as the Monte Carlo results. In addi-
tion, the average number of proteins are almost the same
Case A: MC
Case B: MC
Case A: Approx.
Case B: Approx.
6
for gene i, and si ∈ {ON, OFF} indicates the gene state.
In general, the master equations for multiple gene cases
are very complicated, and it could be difficult to obtain
numerical solutions if the number of genes is large.
However, in our approximation scheme, the effective
interaction is used, and it enables us to reduce the equa-
tions to be solved. Because we consider only the effec-
tive interaction, the probability P (n1, n2, s1, s2) can be
expressed as P (n1, s1)×P (n2, s2). For example, the mas-
ter equation for gene 1 is written as
y
t
i
l
i
b
a
b
o
r
p
0.10
0.08
0.06
0.04
0.02
0.00
0
10
20
30
40
50
60
70
n
FIG. 5: The probability distributions of the number of pro-
teins. Case A: Xeq = 10.0, Xad = 10.0, ω = 0.01. Case B:
Xeq = 50.0, Xad = 50.0, ω = 10.0.
FIG. 6: A schematic illustration of the exclusive switch.
as the Monte Carlo results: for Case A, hni = 1.2 in the
Monte Carlo simulation and hni = 1.3 in the approxima-
tion scheme; for Case B, hni = 19.0 in the Monte Carlo
simulation and hni = 19.0 in the approximation scheme.
IV. EXCLUSIVE SWITCH
Next, we consider a more complicated case, i.e., an
exclusive switch [21 -- 23]. The exclusive switch consists
of two genes, gene 1 and gene 2. The two genes have
overlapping promoter sites, and the binding of one of
the regulatory proteins prevents the binding of the other
regulatory proteins.
Here, the interaction between the binding sites and
proteins is assumed to be monomer interactions. Because
the interaction is not cooperative bindings, the macro-
scopic rate equation gives only one solution [21 -- 23]. Al-
though there is only one solution, it has been shown that
the exclusive switch can play as a switch. In the exclu-
sive switch, stochastic effects make the bistability even
without cooperativity between the regulatory proteins.
In order to study the exclusive switch analytically,
master equations for a joint probability P (n1, n2, s1, s2)
should be constructed; ni ∈ N is the number of proteins
dα′(1)
n
dt
n
dβ(1)
dt
=g(1)[α′(1)
n−1 − α′(1)
n ] + k(1)[(n + 1)α′(1)
n+1 − nα′(1)
n ]
− h(1)α′(1)
n + f (1)β(1)
n ,
= + k(1)[(n + 1)β(1)
n+1 − nβ(1)
n ]
+ h(1)α′(1)
n − f (1)β(1)
n ,
(29)
(30)
where the super-script '(1)' indicates gene 1. Master
equations for gene 2 can be obtained in the similar way.
Using the same discussion in the self-regulating systems
in Sec. III, the generating functions for genes i ∈ {1, 2},
α′(i)(z) and β(i)(z), are derived.
The effective interaction h(1) should be chosen as fol-
lows. The transition of gene 1 from ON state to OFF
state can occur only when the gene 2 is in ON state, and
the effective interaction h(1) includes only a contribution
from the free proteins 2 in ON state. Note that gene 1
does not know whether gene 2 is in ON state or OFF
state, different from the self-regulating system discussed
in Sec. III; in the self-regulating system, the gene knows
the own state. Hence, the evaluation of the effective in-
teraction is slightly different from the self-regulating sys-
tems. The conditional average of the number of free pro-
teins 2 is given by hn(2)iα′ = (∂α′(2)(z)/∂zz=1)/α′(2)(1),
as discussed in Sec. III C. The probability P (2ON), with
which gene 2 is in ON state, is calculated as P (2ON) =
α′(2)(1). Defining hn(2)i′
α′ ≡ hn(2)iα′ P (2ON), the effec-
tive interaction should be written as
h(1) = h(1)hn(2)i′
α′ .
(31)
According to the above discussions, we finally obtain the
following self-consistent equations;
h(2)hn(1)i′
α′ =
h(2)g(1)f (1)(k(1) + f (1))
k(1)(f (1) + h(1)hn(2)i′
α′ )(k(1) + f (1) + h(1)hn(2)i′
α′ )
,
(32)
h(1)hn(2)i′
α′ =
h(1)g(2)f (2)(k(2) + f (2))
k(2)(f (2) + h(2)hn(1)i′
α′ )(k(2) + f (2) + h(2)hn(1)i′
α′ )
.
By solving the above self-consistent equations, we obtain
h(1) and h(2). We here solve them numerically.
(33)
7
(a)
0.03
0.02
0.01
0.00
0
(c)
0.004
0.003
0.002
0.001
0.000
0
25
50
n1
75
100
100
0
50
75
25
n2
25
50
n1
75
100
100
0
50
75
25
n2
(b)
0.03
0.02
0.01
0.00
0
(d)
0.004
0.003
0.002
0.001
0.000
0
25
50
n1
75
100
100
0
50
75
25
n2
25
50
n1
75
100
100
0
50
75
25
n2
Joint probability distributions for the exclusive switch. (a) Monte Carlo results for X (1)
ad =
ad = 25.0, and ω(1) = ω(2) = 0.1. (b) Approximate results. All parameters are the same as (a). (c) Monte Carlo results for
eq = 40.0, X (2)
ad = 20.0, ω(1) = 10.0 and ω(2) = 20.0. (d) Approximate results. All parameters are
ad = 30.0, X (2)
eq = X (2)
eq = 25.0, X (1)
eq = 30.0, X (1)
FIG. 7:
X (2)
X (1)
the same as (c).
We can immediately obtain the probability distribu-
tions for each gene using the approximation scheme. In
order to reconstruct the joint probability distribution
for genes 1 and 2, we need more calculations as fol-
lows. Firstly, we calculate conditional probabilities for
the number of free proteins for gene i (i ∈ {1, 2}) as
α′(i)
n ≡
β(i)
n ≡
n
α′(i)
α′(i)(1)
n
β(i)
β(i)(1)
,
.
(34)
(35)
Secondly, because of the approximation scheme, the joint
probability distribution should be evaluated as
n1−1 α′(2)
P (n1, n2, 1ON, 2ON) = P (1ON, 2ON)α′(1)
P (n1, n2, 1ON, 2OFF) = P (1ON, 2OF)α′(1)
β(2)
n2 ,
n1−1
n1 α′(2)
P (n1, n2, 1OFF, 2ON) = P (1OFF, 2ON) β(1)
n2−1,
n2−1,
(36)
(37)
(38)
where P (1ON, 2ON) is the probability with which gene
1 is in ON state and gene 2 is in ON state, and so on.
Note that α′ means only the number of 'free' proteins; for
monomer interaction cases, the difference between α and
α′ is only one inert protein. In addition, the probability
with which both genes 1 and 2 are in OFF state is zero;
P (1OFF, 2OFF) = 0, because of the exclusive settings.
Taking the exclusive settings
into account,
the
marginal probabilities are calculated as follows:
(39)
(40)
P (1ON, 2ON) + P (1OFF, 2ON) = P (2ON),
P (1ON, 2OFF) = P (2OFF),
P (1OFF, 2ON) = P (1OFF),
P (1ON, 2ON) + P (1ON, 2OFF) = P (1ON),
and then
P (1ON, 2ON) = P (1ON) − P (2OFF),
P (1ON, 2OFF) = P (2OFF),
P (1OFF, 2ON) = P (1OFF),
P (1OFF, 2OFF) = 0.
The marginal probabilities, such as P (1ON), can be evalu-
ated by using the generating function α′(i)(z) and β(i)(z).
Finally, we can construct the joint probability distribu-
tion as
P (n1, n2) =P (n1, n2, 1ON, 2ON)
+ P (n1, n2, 1OFF, 2ON)
+ P (n1, n2, 1ON, 2OFF).
(41)
We here note that the probabilities P (1ON, 2ON), calcu-
lated using the above procedures, may become negative
for some cases; i.e., P (1ON) > P (2OFF) for some choices
of parameters g(i), k(i), h(i) and f (i). In these cases, other
procedures to estimate the joint probabilities are needed.
In the following numerical experiments, only the former
cases (P (1ON) < P (2OFF)) are treated.
eq = X (2)
eq = 25.0, X (1)
eq = 30.0, X (1)
ad = 30.0, X (2)
Figure 7 shows the joint probability distributions. Fig-
ures 7(a) and (c) are Monte Carlo results, and Figs. 7(b)
and (d) are results of the approximation scheme. As
in Sec. III D, we used the rescaled parameters, and set
k(1) = k(2) = 1.
In Figs. 7(a) and (b), we used the
ad = X (2)
parameters X (1)
ad = 25.0,
and ω(1) = ω(2) = 0.1; for (c) and (d), X (1)
eq = 40.0,
X (2)
ad = 20.0, ω(1) = 10.0
and ω(2) = 20.0. The numbers of the Monte Carlo steps
are over 108 for Fig. 7(a), and over 109 for Fig. 7(c).
Although Fig. 7(d) does not show the correlated behav-
ior seen in Fig. 7(c) because correlations between gene
1 and gene 2 are largely neglected in the approximation
scheme, one could say that the approximation scheme
gives qualitatively good results; the characteristics of the
peak structure can be recovered adequately despite the
rough approximation. In Fig. 7(b), the bistability due to
the exclusive settings is recovered well.
8
due to the extension of the basic idea of the effective
interactions, we can naturally apply the approximation
scheme even to the exclusive switch, and the bistability
of the exclusive switch without cooperative interactions
is successfully recovered.
In contrast to the Fokker-Planck or Langevin ap-
the approximation scheme proposed in the
proach,
present paper does not neglect discrete properties of sys-
tems.
In addition, because we can rewritten the joint
probability for all genes as a product of probability dis-
tribution for each gene, the dimensions of the problems
are reduced largely.
The approximation scheme developed in the present
paper would be a very crude one; it cannot treat corre-
lated characteristics between genes. However, approx-
imate analytic expressions are immediately obtained,
and qualitatively good results are given despite the
crude approximation. Since Monte Carlo simulations are
sometimes time-consuming and need high computational
costs, it would be beneficial to study such approxima-
tion scheme in order to obtain qualitative pictures for
the probability distributions. In addition, developments
of analytical treatments would be helpful to gain insights
and understandings for the regulatory systems
V. CONCLUDING REMARKS
ACKNOWLEDGMENTS
In the present paper, we developed the approxima-
tion scheme for gene regulatory systems. We firstly ap-
plied it to self-regulating systems. The approximation
scheme gives qualitatively good results; the characteris-
tics of peak structures can be recovered well. In addition,
The author thank Masaki Sasai for helpful comments
for this manuscript. This work was supported in part
by grant-in-aid for scientific research (Nos. 20115009
and 21740283) from the Ministry of Education, Culture,
Sports, Science and Technology (MEXT), Japan.
[1] C. V. Rao, D. M. Wolf, and A. P. Arkin, Nature 420,
231 (2002).
[2] M. B. Elowitz, A. J. Levine, E. D. Siggia, and P. S. Swain,
Science 297, 1183 (2002).
[12] J. E. M. Hornos, D. Schultz, G. C. P. Innocentini, J.
Wang, A. M. Walczak, J. N. Onuchic, and P. G. Wolynes,
Phys. Rev. E 72, 051907 (2005).
[13] A. M. Walczak, M. Sasai, and P. G. Wolynes, Biophys.
[3] M. Kaern, T. C. Elston, W. J. Blake, and J. J. Collins,
J. 88, 828 (2005).
Nature Rev. Genetics 6, 451 (2005).
[4] T. S. Gardner, C. R. Cantor, and J. J. Collins, Nature
403, 339 (2000).
[14] K.-Y. Kim and J. Wang, PLoS Comp. Biol. 3, e60 (2007).
[15] J. Ohkubo, J. Stat. Mech. P09017 (2007).
[16] V. Shahrezaei and P. S. Swain, Proc. Natl. Acad. Sci
[5] H. Okano, T. J. Kobayashi, H. Tozaki, and H. Kimura,
U.S.A 105, 17256 (2008).
Biophys. J. 95, 1063 (2008).
[17] P. Visco, R. J. Allen, and M. R. Evans, Phys. Rev. E 79,
[6] D. Schultz, J. N. Onuchic, and P. G. Wolynes, J. Chem.
031923 (2009).
Phys. 126, 245102 (2007).
[18] A. M. Walczak and P. G. Wolynes, Biophy. J. 96, 4525
[7] M. Yoda, T. Ushikubo, W. Inoue, and M. Sasai, J. Chem.
(2009).
Phys. 126, 115101 (2007).
[19] J. L. Cherry and F. R. Adler, J. Theor. Biol. 203, 117
[8] J. Hasty, J. Pradines, M. Dolnik, and J. J. Collins, Proc.
(2000).
Natl. Acad. Sci U.S.A 97, 2075 (2000).
[20] P. B. Warren and P. R. ten Wolde, Phys. Rev. Lett. 92,
[9] T. B. Kepler and T. C. Elston, Biophys. J. 81, 3116
128101 (2004).
(2001).
[10] W. Bialek, Adv. Neural Inf. Proc. 13, 103 (2001).
[11] M. Sasai and P.G. Wolynes, Proc. Natl. Acad. Sci U.S.A
100, 2374 (2003).
[21] A. Lipshtat, A. Loinger, N. Q. Balaban, and O. Biham,
Phys. Rev. Lett. 96, 188101 (2006).
[22] A. Loinger, A. Lipshtat, N. Q. Balaban, and O. Biham,
Phys. Rev. E 75, 021904 (2007).
[23] D. Schultz, A. M. Walczak, J. N. Onuchic, and P. G.
Wolynes, Proc. Natl. Acad. Sci U.S.A 105, 19165 (2008).
[24] D. T. Gillespie, J. Phys. Chem. 81, 2340 (1977).
9
|
1807.05937 | 2 | 1807 | 2019-03-25T16:10:21 | Modeling biological systems with an improved fractional Gompertz law | [
"physics.bio-ph",
"math-ph",
"math-ph"
] | The aim of this paper is to provide a fractional generalization of the Gompertz law via a Caputo-like definition of fractional derivative of a function with respect to another function. In particular, we observe that the model presented appears to be substantially different from the other attempt of fractional modifications of this model, since the fractional nature is carried along by the general solution even in its asymptotic behavior for long times. We then validate the presented model by employing it as a reference frame to model three biological systems of peculiar interest for biophysics and environmental engineering, namely: dark fermentation, photofermentation and microalgae biomass growth. | physics.bio-ph | physics |
MODELING BIOLOGICAL SYSTEMS WITH AN IMPROVED
FRACTIONAL GOMPERTZ LAW
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
Abstract. The aim of this paper is to provide a fractional generalization of the Gom-
pertz law via a Caputo-like definition of fractional derivative of a function with respect
to another function. In particular, we observe that the model presented appears to be
substantially different from the other attempt of fractional modifications of this model,
since the fractional nature is carried along by the general solution even in its asymptotic
behavior for long times. We then validate the presented model by employing it as a
reference frame to model three biological systems of peculiar interest for biophysics and
environmental engineering, namely: dark fermentation, photofermentation and microal-
gae biomass growth.
1. Introduction
The Gompertz curve was first introduced in [1] as an empirical model for describing
the human age distribution in a given community, within the analysis of mortality tables.
Then, this distribution has found many applications in biophysics, with particular regard
to the mathematical modeling of tumor growth.
The Gompertz law, from a mathematical perspective, is a Malthusian-type growth
model with a time-dependent exponentially decreasing rate. The stochastic roots of this
model have been carefully addressed by De Lauro et al. in [2] and an interesting general-
ization of this model has been recently presented by Di Crescenzo and Spina in [3].
Applications of fractional calculus [4 -- 7] to counting processes and population modeling
has gained increasing attention over the last few years. This can clearly be inferred from
the extent of the current literature devoted to fractional Poisson processes [8 -- 11], fractional
birth-death processes [12, 13] and fractional Malthusian models [14]. Further, certain
diffusion processes appear to be accurately modeled in terms of fractional kinetic equations
[15]. Such dynamics are known to play a key role in several mechanisms involved in the
life and development of microorganisms. Indeed, as discussed in [16], it is reasonable to
conceive that the physicochemical nature of some relevant biological processes, particularly
those involved in cell growth and enzymatic reactions, will likely show the emergence of
memory effects.
Aside from some interesting modifications involving techniques of ordinary calculus
(see e.g. [17]), a first attempt to consider a generalization of the Gompertz law, by means
of fractional calculus tools, has been proposed by Bolton et al.
in [18]. Besides, very
recently, Almeida and collaborators [14, 19, 20] have shown the peculiar relevance, for
physical applications, of the notion of fractional derivative of a function with respect to
another function.
Date: March 26, 2019.
Key words and phrases. Gompertz growth law; fractional calculus; Mittag-Leffler functions; Fractional
derivative of a function with respect to another function.
In: Comm. Nonlinear Sci. Numer. Simulat. (2019), DOI: 10.1016/j.cnsns.2019.03.024.
1
2
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
Inspired by these investigations, in this paper we suggest a new generalization of the
Gompertz model based on the mathematical approach developed by Almeida in [19].
We compare this new approach with the previous one developed by Bolton et al. [18] and
characterize its main features. Then, we test the proposed model on three distinct biolog-
ical systems that are of particular interest for applications in environmental engineering.
In this regard, let us stress that the key point of this paper is the fractional modification
of the Gompertz law via a Caputo-like definition of fractional derivative of a function with
respect to another function. The experimental validation, even though it involves some
original and unpublished data that might very well be appropriately described in terms
of traditional tools, is therefore regarded as a test of the general features of the proposed
model.
2. Fractional Gompertzian-type models: previous and new results
The seed behind our idea for a generalization of the Gompertz law via fractional calculus
techniques is inspired by a recent study by Bolton et al. [18]. Indeed, in [18] the authors
tested their proposal for a fractional improvement of the aforementioned empirical law by
comparing their theoretical model with an experimental dataset concerning the volume
growth of the Rhabdomyosarcoma tumor in mice.
Let us first recall the key ideas behind the original formulations of both the Gompertz
law and its modification proposed by Bloton and collaborators.
The original formulation of Gompertz's model arises from the following evolution equa-
tion,
1
V
dV
dt
" α ´ β ln
(2.1)
where V " V ptq is the volume of a tumor at time t, V0 " V p0q, while α and β are the so
called kinetic parameters of the tumor (see e.g. [21]). Denoting by y " lnpV {V0q, Eq. (2.1)
reduces to
,
V
V0
Then, going back to the physical variable V ptq, the Gompertz law is obtained in its
(2.2)
whose solution is given by yptq " α
well-known form
(2.3)
dy
dt
" α ´ βy
βp1 ´ e´βtq.
„
´
1 ´ e´βt
α
β
¯
.
V ptq " V0 exp
It is also worth remarking that the Gompertz law is also recovered as a simple Malthu-
sian model (see [3]), namely
(2.4)
where N " Nptq now represents the population size at time t, with an exponentially
time-decreasing rate
" λptq N ,
dN
dt
λptq " αe´βt .
(2.5)
The fractional generalization considered by Bolton et al. in [18] is obtained by replacing
the first order time derivative in (2.2) with a Caputo fractional derivative of order µ P p0, 2q.
According to their analysis, it seems that the physically meaningful range for the fractional
order parameter µ is µ P p0, 1s (in particular, the best agreement between the model and
MODELING BIOLOGICAL SYSTEMS WITH AN IMPROVED FRACTIONAL GOMPERTZ LAW 3
the data is obtained for µ " 0.68). This result can be further justified by the fact that the
solution of this improved model involves the Mittag-Leffler function that has an oscillating
behavior for µ ą 1. Indeed, the general solution of the evolution equation discussed in [18]
is simply given by
Vµptq " V0 exp
,
(2.6)
where
(2.7)
"
α
β
*
r1 ´ Eµp´βtµqs
8ÿ
p´βtµqk
Γpµk ` 1q ,
k"0
Eµp´βtµq "
is the well-known Mittag-Leffler function, that plays a key-role in the theory of fractional
differential equations (see e.g. the recent monograph [22], and [23, 24] for a precise discus-
sion on its numerical evaluation).
Figure 1. Plot of Vµptq in Eq. (2.6) with α " β " V0 " 1 and µ "
0.2, 0.5, 0.6, 1.0 (i.e. the dashed line represents the ordinary Gompertz
law).
In this paper, we provide an alternative analysis of Gompertz's evolution equation based
on the application of fractional Caputo derivatives with respect to another function. This
idea is inspired by the recent studies by Almeida and coauthors [14,19,20], where this new
approach was applied to the study of population growth processes.
Our approach differs substantially from the one considered by Bolton et al. in [18] due
to the different implementation of the fractionalization. Indeed, if we consider the model
equation for the Malthusian model that leads to a Gompertz growth, namely Eq. (2.4)
with the time dependent rate in Eq. (2.5), that reads
(2.8)
eβ t dN
dt
" α N ,
we can simply perform a fractionalization of the latter by introducing the fractional de-
rivative of Nptq with respect to eβt.
t<latexit sha1_base64="CuToXtA5khOi61ixOdNQyKlB0kg=">AAAB9HicbVC7TsNAEFzzDOEVoKQ5kSBRRU4aoIugoQwCk0iJFZ0vm3DK+aG7NSiy8gm0UFEhWv6Hgn/BNi4gYarRzK52drxISUO2/WktLa+srq2XNsqbW9s7u5W9/TsTxlqgI0IV6q7HDSoZoEOSFHYjjdz3FHa8yWXmdx5QGxkGtzSN0PX5OJAjKTil0k2NaoNK1a7bOdgiaRSkCgXag8pXfxiK2MeAhOLG9Bp2RG7CNUmhcFbuxwYjLiZ8jL2UBtxH4yZ51Bk7jg2nkEWomVQsF/H3RsJ9Y6a+l076nO7NvJeJ/3m9mEZnbiKDKCYMRHaIpML8kBFaph0gG0qNRDxLjkwGTHDNiVBLxoVIxTgtpZz20Zj/fpE4zfp53b5uVlsXRTElOIQjOIEGnEILrqANDggYwxM8w4v1aL1ab9b7z+iSVewcwB9YH9+kAZGq</latexit><latexit sha1_base64="CuToXtA5khOi61ixOdNQyKlB0kg=">AAAB9HicbVC7TsNAEFzzDOEVoKQ5kSBRRU4aoIugoQwCk0iJFZ0vm3DK+aG7NSiy8gm0UFEhWv6Hgn/BNi4gYarRzK52drxISUO2/WktLa+srq2XNsqbW9s7u5W9/TsTxlqgI0IV6q7HDSoZoEOSFHYjjdz3FHa8yWXmdx5QGxkGtzSN0PX5OJAjKTil0k2NaoNK1a7bOdgiaRSkCgXag8pXfxiK2MeAhOLG9Bp2RG7CNUmhcFbuxwYjLiZ8jL2UBtxH4yZ51Bk7jg2nkEWomVQsF/H3RsJ9Y6a+l076nO7NvJeJ/3m9mEZnbiKDKCYMRHaIpML8kBFaph0gG0qNRDxLjkwGTHDNiVBLxoVIxTgtpZz20Zj/fpE4zfp53b5uVlsXRTElOIQjOIEGnEILrqANDggYwxM8w4v1aL1ab9b7z+iSVewcwB9YH9+kAZGq</latexit><latexit sha1_base64="CuToXtA5khOi61ixOdNQyKlB0kg=">AAAB9HicbVC7TsNAEFzzDOEVoKQ5kSBRRU4aoIugoQwCk0iJFZ0vm3DK+aG7NSiy8gm0UFEhWv6Hgn/BNi4gYarRzK52drxISUO2/WktLa+srq2XNsqbW9s7u5W9/TsTxlqgI0IV6q7HDSoZoEOSFHYjjdz3FHa8yWXmdx5QGxkGtzSN0PX5OJAjKTil0k2NaoNK1a7bOdgiaRSkCgXag8pXfxiK2MeAhOLG9Bp2RG7CNUmhcFbuxwYjLiZ8jL2UBtxH4yZ51Bk7jg2nkEWomVQsF/H3RsJ9Y6a+l076nO7NvJeJ/3m9mEZnbiKDKCYMRHaIpML8kBFaph0gG0qNRDxLjkwGTHDNiVBLxoVIxTgtpZz20Zj/fpE4zfp53b5uVlsXRTElOIQjOIEGnEILrqANDggYwxM8w4v1aL1ab9b7z+iSVewcwB9YH9+kAZGq</latexit>Vµ(t)<latexit sha1_base64="uOh+CTO7ruhnIXvW5oLRRC7N6CY=">AAAB+nicbVA9TwJBEJ3DL8Qv1NJmI5pgQw4btSPaWGIiHwYuZG8ZcMPu3WV3zoQQfoWtVlbG1lj7Nyz8GfYeB4Wir3p5bybz5vmRkpZc98PJLCwuLa9kV3Nr6xubW/ntnboNYyOwJkIVmqbPLSoZYI0kKWxGBrn2FTb8wcXEb9yhsTIMrmkYoad5P5A9KTgl0k2909YxK9JRJ19wS24K9peUZ6RQOfh6eweAaif/2e6GItYYkFDc2lbZjcgbcUNSKBzn2rHFiIsB72MroQHXaL1RGnjMDmPLKWQRGiYVS0X8uTHi2tqh9pNJzenWznsT8T+vFVPv1BvJIIoJAzE5RFJhesgKI5MmkHWlQSI+SY5MBkxww4nQSMaFSMQ4qSaX9FGe//4vqR2XzkruVdLLOUyRhT3YhyKU4QQqcAlVqIEADffwAI/O2Hlynp2X6WjGme3swi84r9+7Q5bP</latexit><latexit sha1_base64="+dnE1OOVyKQ+u1e5K0ewkZ0vBVc=">AAAB+nicbVC7TgJBFJ3FF4IP1NJmIppgQ3Zt1I5oY4mJPAxsyOxwwQkzu5uZuyRkw1fYamVlbI0/4FdY+BlauywUCp7q5Jx7c889XiiFQdv+sDJLyyura9n1XH5jc2u7sLNbN0GkOdR4IAPd9JgBKXyooUAJzVADU56Ehje4nPiNIWgjAv8GRyG4ivV90ROcYSLd1jttFdESHncKRbtsp6CLxJmRYuXw6+19mP+udgqf7W7AIwU+csmMaTl2iG7MNAouYZxrRwZCxgesD62E+kyBceM08JgeRYZhQEPQVEiaivB7I2bKmJHykknF8M7MexPxP68VYe/MjYUfRgg+nxxCISE9ZLgWSRNAu0IDIpskByp8yplmiKAFZZwnYpRUk0v6cOa/XyS1k/J52b5OerkgU2TJPjkgJeKQU1IhV6RKaoQTRe7JA3m0xtaT9Wy9TEcz1mxnj/yB9foDwqeYSQ==</latexit><latexit sha1_base64="+dnE1OOVyKQ+u1e5K0ewkZ0vBVc=">AAAB+nicbVC7TgJBFJ3FF4IP1NJmIppgQ3Zt1I5oY4mJPAxsyOxwwQkzu5uZuyRkw1fYamVlbI0/4FdY+BlauywUCp7q5Jx7c889XiiFQdv+sDJLyyura9n1XH5jc2u7sLNbN0GkOdR4IAPd9JgBKXyooUAJzVADU56Ehje4nPiNIWgjAv8GRyG4ivV90ROcYSLd1jttFdESHncKRbtsp6CLxJmRYuXw6+19mP+udgqf7W7AIwU+csmMaTl2iG7MNAouYZxrRwZCxgesD62E+kyBceM08JgeRYZhQEPQVEiaivB7I2bKmJHykknF8M7MexPxP68VYe/MjYUfRgg+nxxCISE9ZLgWSRNAu0IDIpskByp8yplmiKAFZZwnYpRUk0v6cOa/XyS1k/J52b5OerkgU2TJPjkgJeKQU1IhV6RKaoQTRe7JA3m0xtaT9Wy9TEcz1mxnj/yB9foDwqeYSQ==</latexit>µ<latexit sha1_base64="vX/u8If68PvrzCh0d1Ox1GlBqjw=">AAAB9HicbVA9TwJBEJ3DL8Qv1FKLjcTEitzZqB3RxhKiJyRwIXvLgBv2PrI7pyGEn2CrlZWxtfavWJj4U7w7LBR81ct7M5k3z4+VNGTbH1ZhYXFpeaW4Wlpb39jcKm/v3Jgo0QJdEalIt3xuUMkQXZKksBVr5IGvsOkPLzK/eYfayCi8plGMXsAHoexLwSmVrjpB0i1X7Kqdg80T54dUavvvjS8AqHfLn51eJJIAQxKKG9N27Ji8MdckhcJJqZMYjLkY8gG2UxryAI03zqNO2GFiOEUsRs2kYrmIvzfGPDBmFPjpZMDp1sx6mfif106of+qNZRgnhKHIDpFUmB8yQsu0A2Q9qZGIZ8mRyZAJrjkRasm4EKmYpKWU0j6c2e/niXtcPavajbSXc5iiCHtwAEfgwAnU4BLq4IKAATzAIzxZ99az9WK9TkcL1s/OLvyB9fYNlKGUbw==</latexit><latexit sha1_base64="h6uJjEnNfwDoBnIkfy+qCF9xxtA=">AAAB9HicbVC7TgJBFJ3FF+ILtdSYicTEiiw2ake0sYToCglsyOxwwQmzj8zc1ZANpaWtVlbG1ppfsfAL/AlnFwoFT3Vyzr255x4vkkKjbX9auYXFpeWV/GphbX1jc6u4vXOrw1hxcHgoQ9X0mAYpAnBQoIRmpID5noSGN7hM/cY9KC3C4AaHEbg+6weiJzhDI123/bhTLNllOwOdJ5UpKVX3x/Xvx4NxrVP8andDHvsQIJdM61bFjtBNmELBJYwK7VhDxPiA9aFlaMB80G6SRR3Ro1gzDGkEigpJMxF+byTM13roe2bSZ3inZ71U/M9rxdg7cxMRRDFCwNNDKCRkhzRXwnQAtCsUILI0OVARUM4UQwQlKOPciLEppWD6qMx+P0+ck/J52a6bXi7IBHmyRw7JMamQU1IlV6RGHMJJnzyRZ/JiPViv1pv1PhnNWdOdXfIH1scPgOKV1Q==</latexit><latexit sha1_base64="h6uJjEnNfwDoBnIkfy+qCF9xxtA=">AAAB9HicbVC7TgJBFJ3FF+ILtdSYicTEiiw2ake0sYToCglsyOxwwQmzj8zc1ZANpaWtVlbG1ppfsfAL/AlnFwoFT3Vyzr255x4vkkKjbX9auYXFpeWV/GphbX1jc6u4vXOrw1hxcHgoQ9X0mAYpAnBQoIRmpID5noSGN7hM/cY9KC3C4AaHEbg+6weiJzhDI123/bhTLNllOwOdJ5UpKVX3x/Xvx4NxrVP8andDHvsQIJdM61bFjtBNmELBJYwK7VhDxPiA9aFlaMB80G6SRR3Ro1gzDGkEigpJMxF+byTM13roe2bSZ3inZ71U/M9rxdg7cxMRRDFCwNNDKCRkhzRXwnQAtCsUILI0OVARUM4UQwQlKOPciLEppWD6qMx+P0+ck/J52a6bXi7IBHmyRw7JMamQU1IlV6RGHMJJnzyRZ/JiPViv1pv1PhnNWdOdXfIH1scPgOKV1Q==</latexit>4
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
µ
Hence, the new model equation becomes
(2.9)
ż
with the initial condition Nµp0q " N0 ą 0, where
C
eβt d
dt
Nµptq " αNµptq, µ P p0, 1s,
´µ
(2.10)
C
eβt d
dt
µ
Nµptq "
1
Γp1 ´ µq
t
e´βτ ´ e´βt
0
β
dNµ
dτ
dτ ,
which, as mentioned above, is a particular case of the Caputo-type fractional derivative of
a function with respect to another function (see e.g. [19] and A for further details). It is
also worth remarking that this approach has been used to treat a Dodson-type fractional
diffusion equation in [25].
It is easy to recognize that Eq. (2.9) is an eigenvalue problem for the Caputo-type
fractional operator defined in Eq. (2.10), whose solution is given by
(2.11)
Nµptq " N0 Eµ
1 ´ e´βt
µ
.
´
„
α
βµ
¯
In A we also provide the solution for a general eigenvalue problem involving a Caputo-type
fractional derivative of a function with respect to another function and we explain how
the general result applies to the case at hand.
Although for both Eq. (2.6) and Eq. (2.11) the classical Gompertz law is recovered for
µ " 1, the main features of the two models clearly differ, as it can be seen in Figure 2.
Indeed, in the first case we have the composition of a fastly increasing function with the
slowly decaying one, the latter being the Mittag-Leffler function, whereas in the second
case the situation is reversed. Indeed, we have that
(2.12)
while
(2.13)
tÑ`8 Nµptq " N0 Eµ
lim
α
βµ
,
tÑ`8 Vµptq " lim
tÑ`8 V1ptq " V0 exppα{βq .
lim
This implies that, asymptotically, the model by Bolton et al. has a classical Gompertz
behavior, while this is not the case for the model introduced in this work. The substantial
difference between these two models is further highlighted by the illuminating plots in
Figure 2.
From a physical perspective, the generalization of the Gompertz law introduced in
Eq. (2.9) has the property of embedding the deterministic time-change (due to the variable
rate) in the definition of the fractional operator. The result of this modification is then
the emergence of memory effects in the population dynamics. Indeed, this can clearly be
inferred from Eq. (2.11) by the peculiar fractional-power law behavior and the distinctive
departure from the typical exponential picture for long times. Besides, the non-local
nature of the operator in Eq. (2.10) is also self-evident in light of Tarasov's criteria [26].
The use of fractional calculus to model bacterial population dynamics is, in our opin-
ion, particularly interesting as it allows us to take into account phenomena like "biomass
adaptation" or "lag-phase", that turn out to be particularly important in batch experi-
ments like those presented in this work. Along this line, it is worth to stress that several
biological studies have shown that bacteria tend to change their features adapting to dif-
ferent environmental conditions. Of course, there are several ways to actively model this
adaptation, from both physical and biological perspectives, yet a common trait of these
MODELING BIOLOGICAL SYSTEMS WITH AN IMPROVED FRACTIONAL GOMPERTZ LAW 5
schemes, particularly when dealing with adaptation strategies of microbial biomass, is the
need to take into account the whole history of the microbial population used for "in-vitro"
experiments. Furthermore, the dynamics of most reactive biological systems is modeled
by means of ordinary differential equations, which however completely neglect any spa-
tial effect. This is usually justified by the assumption that bacterial cells live suspended
in the bulk liquid. Nonetheless, it was recently observed that bacteria tend to organize
themselves in biofilm form [27], which are aggregates of cells enclosed in a self-produced
matrix [28, 29]. Then diffusion phenomena at micro-scale turn out to be of paramount
importance for such systems. In light of this, it was shown that fractional models can
appropriately account for these anomalous kinetics [16].
Figure 2. Comparison, for different values of the parameter µ p0.1 ď
µ ď 1q, between the generalized Gompertz law (2.9) presented here (left:
A1, A2, A3) and the model by Bolton et al. (2.6) (right: B1, B2, B3).
The arrows show how the solutions change as we increase the parameter
µ. Further, the thicker solid line in all panels represents the case µ " 1
(ordinary Gompertz).
6
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
3. Applications and numerical solutions
In the following, we present an application of the model presented in Section 2 to three
different biological systems of major interest for environmental engineering. The first
two cases deal with the biological production of gas (biogas) catalysed by two different
fermentation processes. The third application, instead, concerns the growth of a selected
microalgae biomass in non-limiting nutrient conditions.
Over the last decade, many studies on biogas production through anaerobic fermen-
tative systems have been reported, see e.g. [30, 31]. Due to the production of a high
energy content gaseous product, such as hydrogen, accompanied solely by water during
its combustion, Dark Fermentation (DF) is widely regarded as one of the most promis-
ing anaerobic biological process for engineering purposes [32]. This process is usually
performed by multispecies anaerobic consortia, which provide the conversion of organic
substrates, e.g. the organic fraction of municipal solid wastes, in a biologically derived
fuel that can be used to produce energy. The wide range of organic (solid-liquid) waste
compounds that can act as feedstock for the DF has led this process to become one of
the key technique for renewable energy production in both laboratories and pilot scale
reactors.
Similar results can be achieved by employing Photofermentative bacteria, such as Rodobac-
ter Sphaeroides (Second application). These species are able to catalyze the conversion
of different organic compounds by using light as a supplementary energy source [33, 34].
In this case, the organic soluble compounds, which are usually derived from a clarified
wastewater, are involved in biochemical reactions that lead to the accumulation of poly-
hydroxybutyrate (PHB) and the contextual production of hydrogen. PHBs are value-
added biochemical compounds as they can be easily extracted from the bacterial cells and
processed to produce bio-plastic materials [35]. Besides, the hydrogen produced in these
process can be used as a renewable energy source, as it ensure a high energy recovery from
organic wastes.
Finally, the third application of the model introduced in Section 2 deals with the descrip-
tion of the microalgae biomass growth. These unicellular micro-organisms can be used in
the wastewater treatment field for the removal of nutrients, as an alternative treatment to
the traditional, energy demanding, aerobic systems [36]. Microalgae can assimilate a large
amount of inorganic nitrogen and phosphorous in autotrophic conditions by employing
carbon dioxide as inorganic carbon source and mitigating greenhouse gas emissions with
the contextual production of oxygen. Moreover, these micro-organisms, depending on the
conditions at hand, can shift their metabolism from autotrophic to heterotrophic or, even,
assimilate organic compounds to produce hydrogen [37]. In addition, many microalgae
species are able to accumulate value-added compounds during their growth, usually due
to the high protein content of their cell membrane, which can be later extracted or used
for biofuel production.
The complete description of the lab-scale procedures adopted for the experimental data
acquisition have been reported in the Supplementary Materials (B), whereas the model
application and validation is reported in the following Section.
3.1. Model validation. The modified Gompertz law proposed in this work has been
applied to the three experimental cases described above. The calibrated parameters used in
numerical simulation are reported in Table 1, whereas a comparison between experimental
data and numerical simulations for each case is reported in Figure 3.
(3.1)
(3.2)
N
i"1 pPi ´ Oiq2
dř
nRMSE " 1
ř
ř
O
i"1 pPi ´ Oiq2
IoA " 1 ´
`
ř
ř
i"1 pP1
i ` O1
`
ř
i"1 pPi ´ Oiq2
Oi ´ O
2 ´
Oi ´ O
N
i"1
iq2 ,
N
i"1
N
,
N
N
N
MODELING BIOLOGICAL SYSTEMS WITH AN IMPROVED FRACTIONAL GOMPERTZ LAW 7
Table 1. Model parameters
Parameter
α
β
V0
µ
A
1.0965
0.180
0.020
0.64
B
1.5200
0.275
0.001
0.70
C
2.1350
0.350
0.010
0.85
The quantitative comparison between observed and predicted data has been carried out
via three standard methods for model validation [38,39], namely the normalized root mean
square error (nRMSE), the index of agreement (IoA), and the modeling efficiency (ME).
Specifically,
ME "
(3.3)
where Pi and Oi denote the predicted and observed values, i " 1, . . . , N , N is the number of
i " Oi´ O.
observed/predicted values, O is the mean of observed data, and P1
i " Pi´ O; O1
2
;
The results of the validation procedure are then depicted in Figure 3. Besides, a compar-
ison between experimental data and numerical simulations is also summarized in Table 2.
Table 2. Results of model validation
Experiment
Dark Fermentation
Photofermentation
Microalgae
nRMSE IoA
0.0374
0.0792
0.1672
0.9988
0.9961
0.9918
ME
0.9952
0.9835
0.9693
4. Conclusion
In this paper we introduced a mathematical model that generalizes the well-known
Gompertz law of population dynamics by including fractional features in its constitutive
(ordinary) differential equation. Specifically, this procedure was performed by replacing
the left-hand side in Eq. (2.8) with the Caputo-type derivative of Nptq with respect to
expp´β tq. The resulting model is then governed by a purely fractional differential equation
whose solution of the corresponding initial value problem (2.11) shows that the fractional
nature of the described theoretical set-up is carried along for the whole history of the
process. Indeed, one of the key differences between the behavior of (2.11) and the model
by Bolton et al (2.6) is that for the latter the contribution of the fractional feature get
slowly turned off at late time, ultimately stabilizing the evolution to an asymptotic value
(2.13) which does not depend on the fractional parameter µ. Whereas, the new growth law
featured in (2.11) is still strongly dependent on the fractionalization procedure as shown
8
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
Figure 3. Comparison between experimental data and numerical simula-
tions for the three experimental cases: (A) Dark Fermentation, (B) Photo
fermentation, (C) Algae.
in (2.12), that underlines an enhanced capability of this model to fit a broader class of
experimental behaviors.
In order to further strengthen our arguments, we have tested the performance of the
presented model in three very peculiar and distinct frameworks, namely: two experiments
regarding the biogas production catalyzed by either DF (B.1) or PF (B.2), and one exper-
iment featuring microalgae biomass growth (B.3). The results concerning the calibration
and numerical simulations are reported in Section 3.1, together with some illuminating
plots, see Figure 3.
Lastly, a comment is in order concerning the general reasoning behind the very sig-
nificance of the whole fractionalization process. The underlying motivation for this work
0510152025303540t [hour]05101520H2 [mmol]A051015t [day]051015SS [gr L-1]C051015202530t [day]00.511.5H2 [mol/molC6H12O6]BMODELING BIOLOGICAL SYSTEMS WITH AN IMPROVED FRACTIONAL GOMPERTZ LAW 9
was to analyze what kind of information could be retrieved by the introduction of the
newly developed non-local operator that, as stressed above, clearly improves on previous
attempts at generalizing the Gompertz law. Then, we decided to test the effectiveness
of these modified paradigm upon experimental data which are usually tackled with the
standard Gompertz law. Of course, it is well known that not all biological systems follow
the usual Gompertz law, so this is per se a valid justification for looking for evolution
equations whose solution might even depart sensibly form the latter. However, the typical
issue that is commonly raised when dealing with fractional operators is connected with
their vindication starting from first principles. This is usually a very difficult matter to
address since, most of the time, it requires a case-dependent bottom-up construction of the
theory (see e.g. [15, 40 -- 43] and references therein). Hence, we leave this matter regarding
the hereby presented modified Gompertz law for a future study.
Appendix A. Preliminaries about Caputo fractional derivative of a
function with respect to another function
Fractional derivatives of a function with respect to another function have been part of
the fractional calculus literature at least since the publication of the classical monograph
by Kilbas et al. [44] (see Section 2.5) and appear to have found a new life tanks to some
recent works by Almeida [19].
In this Appendix, we briefly recall the main definitions and properties of these opera-
tors. Nonetheless, this section cannot represent an exhaustive description of these pecu-
liar operators, hence we invite the interested reader to refer to [19] for a more complete
discussion. We also observe that this approach has found some interesting applications
in [14, 20, 25, 45, 46].
Definition 1. Let ν ą 0, I " pa, bq be an interval such that ´8 ď a ă b ď `8,
gptq P L1pIq and fptq P C1pIq strictly increasing function for all t P I. Then, the fractional
integral of a function gptq with respect to another function fptq is given by
(A.1)
a` gptq :" 1
I ν,f
Γpνq
t
f1pτqrfptq ´ fpτqsν´1gpτqdτ.
ż
a
By inspection one can immediately notice that for fptq " tα{β we recover the Erd´elyi-
Kober fractional integral (see e.g. [47, 48] for some physical applications), whereas if we
set fptq " ln t, this integral can be recasted in the Hadamard fractional integral. Finally,
for fptq " t one trivially recovers the Riemann-Liouville fractional integral.
Now, given Definition 1 one can introduce a corresponding regularized left-inverse op-
erator of given by,
Definition 2. Let ν ą 0, n " rνs, I " pa, bq be an interval such that ´8 ď a ă b ď `8,
gptq P CnpIq and fptq P C1pIq strictly increasing function for all t P I. Then, the Caputo
derivative of the function gptq with respect to a function fptq is given by
(A.2)
C
1
f1ptq d
dt
ν
gptq :" I n´ν,f
a`
1
f1ptq d
dt
n
gptq .
For the latter it is easy to infer that,
Proposition 1. Let ν ą 0, n " rνs, I " pa, bq be an interval such that ´8 ď a ă b ď `8
and fptq P C1pIq strictly increasing function for all t P I. If gptq " rfptq´ fpa`qsβ´1 with
10
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
β ą 1, then
ν
C
1
f1ptq d
(A.3)
Proposition 2. Let ν ą 0, n " rνs, I " pa, bq be an interval such that ´8 ď a ă b ď `8
and fptq P C1pIq strictly increasing function for all t P I. If gptq " Eνrλpfptq ´ fpa`qqνs
with λ P R, then
gptq " Γpβq
Γpβ ´ νqrfptq ´ fpa`qsβ´ν´1 .
dt
(A.4)
C
1
f1ptq d
dt
ν
gptq " λ gptq .
„
α
βµp1 ´ e´βtq
.
The operator appearing in the generalized Gompertz equation (2.9) is obtained from
the general definition (A.2) by setting a " 0` and choosing fptq such that f1ptq " e´βt.
Hence, from 2 we have that the eigenfunction of the operator (2.10) is given by
(A.5)
Nµptq " Eµ
Appendix B. Supplementary Materials
In this Section we report a detailed description of the three experimental setups for the
model validation discussed in the manuscript.
B.1. Dark Fermentation experiments. DF experiments have been inoculated with a
pre-treated full-scale anaerobic digestion sludge. The thermal pre-treatment leads to the
inhibition of methanogenic bacteria with consequent hydrogen accumulation during the
fermentation process.
Figure 4. Photofermentative experiment
According to [31], the collected sludge was retained in a lab-scale oven for one hour at
105 C. The anaerobic sludge was characterized in terms of Total Solids (TS), Volatile
Solids (VS) and pH, according to the standard methods [?]. The results showed TS
and VS content and a pH value of 28.75 g{L, 18.90 g{L and 7.7 g{L respectively. The
inoculum, 200 mL, was added into an airtight 1000 mL transparent glass bottle (Simax,
MODELING BIOLOGICAL SYSTEMS WITH AN IMPROVED FRACTIONAL GOMPERTZ LAW 11
Czech Republic) and 100 mL of a glucose solution (17.71 g{L) was used as substrate. The
final volume of 600 mL was reached by using tap water.
The reactors were connected to an automatic system for data acquisition and contin-
uous pH control through NaOH addition as in [32]. Different pH set-point values in the
range 5.5 ´ 7.0 were adopted in the experiments. The substrate to inoculum ratio F{M
(Food/Microorganisms) was fixed at 1.5 (g COD1 substrate/g VS inoculum).
Each batch bioreactor was immersed into a thermostatic bath at p35 1q C to ensure
mesophilic conditions. This procedure was applied to three different DF reactors and,
before closing the bottles, thirty minutes of nitrogen purge was carried out under anaerobic
hood. The plastic cap of each reactor was equipped with sampling tubing. The liquid and
gaseous samples were withdrawn by connecting the plastic cups to a gas measurement
system (eudiometer) and a syringe, respectively.
The total amount of the produced biogas was determined by water displacement. Biogas
was characterized by a Varian Star 3400 gas chromatograph equipped with ShinCarbon
ST 80/100 column and a thermal conductivity detector. Argon was used as carrier gas
with 1.4 bar front and rear end pressure. The Organic Acids (OA) concentrations were
quantified by high pressure liquid chromatography (HPLC) (Dionex LC 25 Chromatogra-
phy Oven) equipped with a Synergi 4u Hydro RP 80A (size 250 4.60 mm) column and
UV detector (Dionex AD25 Absorbance Detector). The cumulative normalized hydrogen
production over time is reported in Figure 3A.
B.2. Photofermentative experiments. An adapted mixed culture of Purple Non-Sulphur
Bacteria (PNSB) isolated from the Averno Lake (Naples, Italy) and enriched in a lab-scale
reactor under continuous illumination was tested. In particular, the batch experiments
were carried out in triplicate by using 500 mL borosilicate glass bottles GL 45 (Shott Du-
ran, Germany) with a 400 mL working volume. The caps of each reactor were equipped
with thin tubing on the top for sampling and gas extraction as in [33]. The light was
continuously provided through fluorescent lamps which ensure constant illumination of
4000 lx. The stirring conditions were fixed to 300 rpm through IKA RT 5 stirrer stations.
The headspace of the reactors was flushed with argon gas for 20 min to avoid nitroge-
nase inhibition due to high nitrogen content [34]. The Photo Fermentation (PF) reactors
were fed with a synthetic culture medium reproducing the characteristics of a real Dark
Fermentation effluent [35]. The experiments were executed at room temperature (25 C).
The sampling time for gas and liquid analysis varied from two to five days, depending on
the PNSB growth phase and their activity. Hydrogen production was quantified through
water displacement and gas chromatography by using a 1 L column fulfilled with an acidic
solution (1.5 % HCl) to avoid biogas solubilisation. The hydrogen production was later
calculated by considering the total biogas composition under normal conditions. The final
hydrogen production was later normalized with respect to the organic load of glucose
supplemented to the reactors (Figure 3B).
B.3. Microalgae Experiments. The experimental set-up was similar to the PF tests.
The Chlorella pyrenoidosa inoculum was provided by the "Algae Biology Laboratory" of
the University of Naples "Federico II" (Italy). All the tests were carried out at room
temperature in batch conditions by using 500 mL borosilicate glass bottles GL 45 (Shott
Duran, Germany), constantly mixed at 200 rpm. Continuous illumination was provided
by using fluorescent lamps (average light intensity of 4000 lx).
1Where COD (Chemical Oxygen Demand) is a laboratory experimental measure of the organic com-
pounds in a solid or liquid matrix.
12
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
Figure 5. Algae experiment
A modified Bold's Basal Medium [36] with nitrate concentration around 180 mgN O´
3 {L
was used for microalgae enrichment. Autotrophic growth was stimulated by the addition of
inorganic carbon (N aHCO3) until reach the C{N (Carbon/Nitrogen) ratio of 15. Biomass
growth was monitored trough OD 680 and a correlation curve was used to achieve the
specific dry weight trend over time (Figure 3C).
Acknowledgments
This work has been carried out in the framework of the activities of the National Group
of Mathematical Physics (GNFM, INdAM). Moreover, the work of L.F. and A.G. is par-
tially supported by GNFM/INdAM Young Researchers Project 2017. Finally, L.F. and
V.L. would also like to acknowledge the support of the project VOLAC - Valorization of
OLive oil wastes for sustainable production of biocide-free Antibiofilm Compounds, funded
by CARIPLO foundation.
The experiments reported in this work have been performed by Dr Vincenzo Luongo at
the Laboratory of Analysis and Environmental Research (LARA) of The Department of
Civil, Architectural and Environmental Engineering (DICEA) of the University of Naples
"Federico II ".
References
[1] Gompertz B. On the nature of the function expressive of the law of human mortality, and on a new
mode of determining the value of life contingencies. Philosophical Transactions of the Royal Society of
London, 1825, 155, 513 -- 583.
[2] De Lauro E, De Martino S, De Siena S, Giorno V. Stochastic roots of growth phenomena. Physica A,
2014, 401, 207 -- 213.
[3] Di Crescenzo A, Spina S. Analysis of a growth model inspired by Gompertz and Korf laws, and an
analogous birth-death process. Mathematical biosciences, 2016, 282, 121 -- 134.
[4] Giusti A. A comment on some new definitions of fractional derivative. Nonlinear Dynamics, 2018, 93,
1757 -- 1763.
[5] Gorenflo R, Mainardi F. Fractional Calculus: Integral and Differential Equations of Fractional Order.
In Fractals and Fractional Calculus in Continuum Mechanics; Carpinteri, A., Mainardi, F., Eds.; Springer:
New York, NY, USA; Wien, Austria, 1997.
MODELING BIOLOGICAL SYSTEMS WITH AN IMPROVED FRACTIONAL GOMPERTZ LAW 13
[6] Pirozzi E. Colored noise and a stochastic fractional model for correlated inputs and adaptation in
neuronal firing. Biological cybernetics, 2018, 112, 25 -- 39.
[7] Mainardi F. Fractional Calculus: Some basic problems in continuum and statistical mechanics. In
Fractals and Fractional Calculus in Continuum Mechanics; Carpinteri, A., Mainardi, F., Ed.; Springer:
New York, NY, USA; Wien, Austria, 1997.
[8] Beghin L, Orsingher E. Fractional Poisson processes and related planar random motions. Electronic
Journal of Probability, 2009, 14, 1790 -- 1826.
[9] Konno H, Yoshiyasu T. Stochastic modeling for neural spiking events based on fractional superstatistical
Poisson process. AIP Advances, 2018, 8, 015118.
[10] Laskin N. Fractional Poisson process. Communications in Nonlinear Science and Numerical Simula-
tion, 2003, 8, 201 -- 213.
[11] Mainardi F, Gorenflo R, Scalas E. A fractional generalization of the Poisson processes. Vietnam
Journal of Mathematics, 2013, 32, 53 -- 64.
[12] Orsingher E, Polito F. On a fractional linear birth-death process. Bernoulli, 2011, 17, 114 -- 137.
[13] Orsingher E, Ricciuti C, Toaldo B. Population models at stochastic times. Advances in Applied Prob-
ability, 2016, 48, 481 -- 498.
[14] Almeida R. What is the best fractional derivative to fit data?. Applicable Analysis and Discrete Math-
ematics, 2017, 11, 358 -- 368.
[15] Metzler R, Klafter J. The random walk's guide to anomalous diffusion: a fractional dynamics approach.
Physics reports, 2000, 339, 1 -- 77.
[16] Toledo-Hernandez R, Rico-Ramirez V, Iglesias-Silva G, Diwekar U. A fractional calculus approach to
the dynamic optimization of biological reactive systems. Part I: Fractional models for biological reactions.
Chemical Engineering Science, 2014, 17, 217 -- 228.
[17] Tjørve K, Tjørve E. The use of Gompertz models in growth analyses, and new Gompertz-model
approach: An addition to the Unified-Richards family. PloS one, 2017, 12, e0178691.
[18] Bolton L, Cloot A, Schoombie S, Slabbert J. A proposed fractional-order Gompertz model and its
application to tumour growth data. Mathematical Medicine and Biology, 2014, 32, 187 -- 207.
[19] Almeida R. A Caputo fractional derivative of a function with respect to another function. Communi-
cations in Nonlinear Science and Numerical Simulation, 2017, 44, 460 -- 481.
[20] Almeida R, Malinowska A, Monteiro M. Fractional differential equations with a Caputo derivative
with respect to a kernel function and their applications. Mathematical Methods in the Applied Sciences,
2018, 41, 336 -- 352.
[21] Wheldon T. Mathematical Models in Cancer Research; Bristol: Adam Hilger, IOP Publishing Ltd.,
1988.
[22] Gorenflo R, Kilbas A, Mainardi F, Rogosin S. Mittag-Leffler functions, related topics and applications;
Berlin: Springer, 2014.
[23] R. Garrappa, M. Popolizio, Evaluation of generalized Mittag-Leffler functions on the real line, Ad-
vances in Computational Mathematics, 2013, 39, 205 -- 225.
[24] R. Garrappa, Numerical Evaluation of two and three parameter Mittag-Leffler functions, SIAM Jour-
nal of Numerical Analysis, 2015, 53, 1350 -- 1369.
[25] Garra R, Giusti A, Mainardi F. The fractional Dodson diffusion equation:
a new approach.
Ricerche di Matematica, 2018, 67(2), 899 -- 909.
[26] Tarasov, V. No nonlocality. No fractional derivative. Communications in Nonlinear Science and Nu-
merical Simulation, 2018, 62, 157 -- 163.
[27] Flemming HC, Wingender J, Szewzyk U, Steinberg P, Rice SA, Kjelleberg S. Biofilms: an emergent
form of bacterial life. Nature Reviews Microbiology, 2016, 14, 563 -- 575.
[28] Mattei MR, Frunzo L, D Acunto B, Pechaud Y, Pirozzi F, Esposito G. Continuum and discrete
approach in modeling biofilm development and structure: a review. Journal of Mathematical Biology,
2018, 76(4), 945 -- 1003.
[29] D'Acunto B, Frunzo L, Klapper I, Mattei MR, Stoodley P. Mathematical modeling of dispersal phe-
nomenon in biofilms. Mathematical Biosciences, 2019, 307, 70 -- 87.
[30] De Gioannis G, Muntoni A, Polettini A, Pomi R. A review of dark fermentative hydrogen production
from biodegradable municipal waste fractions. Waste management, 2013, 33, 1345 -- 1361.
[31] Ghimire A, Frunzo L, Pirozzi F, et al. A review on dark fermentative biohydrogen production from
organic biomass: process parameters and use of by-products. Applied Energy, 2015, 144, 73 -- 95.
[32] De Gioannis G, Friargiu M, Massi E, et al. Biohydrogen production from dark fermentation of cheese
whey: Influence of pH. International Journal of Hydrogen Energy, 2014, 39(36), 20930 -- 20941.
14
LUIGI FRUNZO1, ROBERTO GARRA2, ANDREA GIUSTI3, AND VINCENZO LUONGO4
[33] Ghimire A, Valentino S, Frunzo L, et al. Concomitant biohydrogen and poly-β-hydroxybutyrate pro-
duction from dark fermentation effluents by adapted Rhodobacter sphaeroides and mixed photofermen-
tative cultures. Bioresource Technology, 2016, 217, 157 -- 164.
[34] Trchounian K, Sawers R, Trchounian A. Improving biohydrogen productivity by microbial dark-and
photo-fermentations: Novel data and future approaches. Renewable and Sustainable Energy Reviews,
2017, 80, 1201 -- 1216.
[35] Luongo V, Ghimire A, Frunzo L, et al. Photofermentative production of hydrogen and poly-β-
hydroxybutyrate from dark fermentation products. Bioresource Technology, 2017, 228, 171 -- 175.
[36] Iasimone F, De Felice V, Panico A, Pirozzi F. Experimental study for the reduction of CO 2 emissions
in wastewater treatment plant using microalgal cultivation. Journal of CO2 Utilization, 2017, 22, 1 -- 8.
[37] Osorio J, Luongo V, Del Mondo A, et al. Nutrient removal from high strength nitrate containing
industrial wastewater using Chlorella sp. strain ACUF 802, Annals of Microbiology, 2018, 68, 899 -- 913.
[38] Janssen P, Heuberger P. Calibration of process-oriented models. Ecological Modelling, 1995, 83(1-2),
55 -- 66.
[39] Frunzo L, Esposito G, Pirozzi F, Lens P. Dynamic mathematical modeling of sulfate reducing gas-lift
reactors. Process Biochemistry, 2012, 47, 2172 -- 2182.
[40] Pagnini G. Short note on the emergence of fractional kinetics. Physica A, 2014, 409, 29 -- 34.
[41] Giusti A, Mainardi F. A dynamic viscoelastic analogy for fluid-filled elastic tubes. Meccanica, 2016,
51, 2321 -- 2330.
[42] Molina-Garc´ıa D, et al. Fractional kinetics emerging from ergodicity breaking in random media. Phys-
ical Review E, 2016, 94, 052147.
[43] Colombaro I, Giusti A, Mainardi, F. On the propagation of transient waves in a viscoelastic Bessel
medium. Zeitschrift fur Angewandte Mathematik und Physik, 2017, 68, 62 -- 74.
[44] Kilbas A, Srivastava H, Trujillo J. Theory and applications of fractional differential equations; Vol.
204. Elsevier Science Limited, 2006.
[45] Colombaro I, Garra R, Giusti A, Mainardi F. Scott-Blair models with time-varying viscosity. Ap-
plied Mathematics Letters, 2018, 86, 57 -- 63.
[46] Jleli M, O'Regan D, Samet B. Some fractional integral inequalities involving m-convex functions.
Aequationes Mathematicae, 2017, 91, 479 -- 490.
[47] Pagnini G. Erd´elyi-Kober fractional diffusion. Fractional Calculus and Applied Analysis, 2012, 15,
117 -- 127.
[48] R. Garra, A. Giusti, F. Mainardi, G. Pagnini, Fractional relaxation with time-varying coefficient,
Fractional Calculus and Applied Analysis, 2014, 17, 424 -- 439.
1 University of Naples "Federico II", Department of Mathematics and Applications, via
Cintia 1, 80126, Naples, ITALY.
E-mail address: [email protected]
2 Department of Statistical Sciences, Sapienza, University of Rome. P.le Aldo Moro, 5,
Rome, ITALY
E-mail address: [email protected]
3 Department of Physics & Astronomy, University of Bologna and INFN. Via Irnerio 46,
Bologna, ITALY and Arnold Sommerfeld Center, Ludwig-Maximilians-Universitat, There-
sienstrasse 37, 80333 Munchen, GERMANY.
E-mail address: [email protected]
4 University of Naples "Federico II", Department of Mathematics and Applications, via
Cintia 1, 80126, Naples, ITALY.
E-mail address: [email protected]
|
1803.00522 | 1 | 1803 | 2018-03-01T17:38:01 | Modelling of surfactant-driven front instabilities in spreading bacterial colonies | [
"physics.bio-ph",
"cond-mat.soft"
] | The spreading of bacterial colonies at solid-air interfaces is determined by the physico-chemical properties of the involved interfaces. The production of surfactant molecules by bacteria is a widespread strategy that allows the colony to efficiently expand over the substrate. On the one hand, surfactant molecules lower the surface tension of the colony, effectively increasing the wettability of the substrate, which facilitates spreading. On the other hand, gradients in the surface concentration of surfactant molecules result in Marangoni flows that drive spreading. These flows may cause an instability of the circular colony shape and the subsequent formation of fingers. In this work, we study the effect of bacterial surfactant production and substrate wettability on colony growth and shape within the framework of a hydrodynamic thin film model. We show that variations in the wettability and surfactant production are sufficient to reproduce four different types of colony growth, which have been described in the literature, namely, arrested and continuous spreading of circular colonies, slightly modulated front lines and the formation of pronounced fingers. | physics.bio-ph | physics | Modelling of surfactant-driven front instabilities in spreading bacterial colonies
Sarah Trinschek,1, 2 Karin John,2 and Uwe Thiele1, 3, 4, ∗
1Institut für Theoretische Physik, Westfälische Wilhelms-Universität Münster,
Wilhelm Klemm Str. 9, 48149 Münster, Germany
2Université Grenoble-Alpes, CNRS, Laboratoire Interdisciplinaire de Physique, 38000 Grenoble, France
3Center of Nonlinear Science (CeNoS), Westfälische Wilhelms-Universität Münster, Corrensstr. 2, 48149 Münster, Germany
4Center for Multiscale Theory and Computation (CMTC),
Westfälische Wilhelms-Universität, Corrensstr. 40, 48149 Münster, Germany
8
1
0
2
r
a
M
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
2
2
5
0
0
.
3
0
8
1
:
v
i
X
r
a
The spreading of bacterial colonies at solid-air interfaces is determined by the physico-chemical
properties of the involved interfaces. The production of surfactant molecules by bacteria is a
widespread strategy that allows the colony to efficiently expand over the substrate. On the one
hand, surfactant molecules lower the surface tension of the colony, effectively increasing the wet-
tability of the substrate, which facilitates spreading. On the other hand, gradients in the surface
concentration of surfactant molecules result in Marangoni flows that drive spreading. These flows
may cause an instability of the circular colony shape and the subsequent formation of fingers. In
this work, we study the effect of bacterial surfactant production and substrate wettability on colony
growth and shape within the framework of a hydrodynamic thin film model. We show that variations
in the wettability and surfactant production are sufficient to reproduce four different types of colony
growth, which have been described in the literature, namely, arrested and continuous spreading of
circular colonies, slightly modulated front lines and the formation of pronounced fingers.
I.
INTRODUCTION
Bacteria are able to colonize solid-air interfaces by the
formation of dense colonies.[1] After the attachment of in-
dividual bacteria to the surface, they proliferate and a dense
colony starts to expand laterally over the surface. In many
cases, the spreading is not driven by the active mobility
of individual bacteria but rather by growth processes and
passive flows that result from the physico-chemical proper-
ties of the bacterial film and the substrate.[2, 3] One well
studied example is the osmotic spreading of biofilms, where
the bacteria secrete an extracellular matrix that acts as an
osmolyte and triggers the influx of nutrient-rich water from
the underlying moist agar substrate into the colony, which
subsequently swells and spreads out.[3–7]
Another physical effect that plays a role in the spreading
of bacterial colonies at solid-air interfaces are wetting phe-
nomena, which govern the motion of the three-phase con-
tact line between the colony, the underlying agar substrate,
and the surrounding air. For many bacterial strains, the
molecules which are involved in the quorum sensing mech-
anism (which allows for a cell-cell communication) have
been found to play a double role. Beside their signalling
function, they act as bio-surfactants (small molecules which
adsorb to surfaces, thereby lowering the surface tension) at
physiologically relevant concentrations.[8, 9] Measurements
of surface tension and contact angle [10, 11] indicate that
bio-surfactants promote the spreading of bacterial colonies
by improving wettability. Additionally, gradients in surfac-
tant concentration at the edges of the colony give rise to
so-called Marangoni fluxes which further drive cooperative
spreading.[3, 12–14]
For Rhinozobium etli, genetic knock-out experiments[15]
show that AHL (N-acyl-homoserine lactone) molecules are
∗ [email protected]
crucial for an efficient swarming of the colony. The experi-
mentally observed spreading speeds and colony shapes are
consistent with those estimated from a spreading driven by
Marangoni forces. Growth measurements verify that for
Paenibacillus dendritiformis colonies, the spreading veloc-
ity indeed depends on the surfactant concentration but not
on the individual bacterial motion.[16] Genetic and physico-
chemical experiments [3, 12, 14, 17, 18] show that also
in Bacillus subtilis and Pseudomonas aeruginosa colonies,
the surface tension gradient induced by the respective bio-
surfactants surfactin and rhamnolipids is an important
driver of colony expansion. Further support of this the-
ory comes from the demonstration that swarming can be
inhibited by the rhamnolipid production of nearby colonies
[19] as well as by the addition of purified rhamnolipids to
the agar substrate [14] as both suppress the necessary gra-
dients in surface tension. Besides enhancing the spreading
speed, Marangoni fluxes may also be responsible for the
striking dendritic or finger-like colony patterns observed in
swarming experiments. Surfactant-producing Pseusomona
aeruginosa colonies spread outwards and form pronounced
fingers whereas a mutant strain deficient in surfactant pro-
duction can not expand and is arrested in a small circular
shape.[12, 14]
In the surfactant-assisted spreading of liquid drops (see
20 for a review), Marangoni fluxes are known to give rise to
a fingering front instability as first observed experimentally
in 21 and subsequently confirmed and studied in detail, e.g.,
in 22–26. Numerical time simulations and transient growth
analysis of hydrodynamic thin-film models show the pres-
ence of the instability for films covered by insoluble sur-
factants [27, 28], but are also extended to soluble surfac-
tants with sorption kinetics [29] and micelle formation.[30]
In the context of surfactant-mediated spreading of bacterial
colonies, similar thin film models are successfully applied
to study the movement of a Bacillus subtilis biofilm up a
wall on waves of surfactant[18] or bacterial swarming in
2
fore introducing the bioactive terms.
A. Passive part of the model
We consider a thin film of height h(x, y, t) which is cov-
ered by an insoluble surfactant of (area-)density Γ(x, y, t).
The description of the passive part of the model is based
on the free energy functional
F [h, Γ] =
[fw(h) + fs (Γ) ζ] dx
with
(cid:112)
1 + ∇h2 ≈ 1 + 1
2∇h2 .
ζ =
(1)
(2)
F [h, Γ] in (1) contains the wetting energy fw(h) and the lo-
cal free energy of the surfactant-covered free surface fs(Γ).
Here, ds = ζdx is the surface element of the curved liquid
surface and dx is the surface element of the euclidean flat
substrate plane as depicted in Fig. 1. A common choice
for the wetting energy is[40]
(cid:18)
(cid:19)
fw(h) = A
−
1
2h2 +
h3
a
5h5
,
(3)
which combines destabilizing long-range van-der-Waals and
stabilizing short-range interactions. It describes a partially
wetting fluid, i.e. a macroscopic drop sitting on a stable
adsorption layer of thickness ha.
Assuming relatively low densities of surfactant, the contri-
bution of a non-interacting surfactant to the energy of the
interface corresponds to an entropic term
fs(Γ) = γ +
kT
a2 Γ [log(Γ) − 1]
(4)
that results in the usual linear equation of state. Here,
γ denotes the surface tension, kT the thermal energy and
a2 is the effective area of the surfactant molecules on the
interface. In order to write evolution equations for the film
height and the surfactant concentration in the formulation
as a gradient dynamics [32, 41], it is necessary to introduce
the projection of the area density onto the flat surface of
the substrate
Γ(x, y, t) = ζΓ(x, y, t) .
(5)
The free energy functional (1) can now be used to write
evolution equations for h(x, y, t) and Γ(x, y, t)
∂th = ∇ ·
∂t Γ = ∇ ·
(cid:20)
(cid:20)
(cid:18)Qhh QΓh
QhΓ QΓΓ
Q =
(cid:21)
(cid:21)
(cid:33)
Qhh∇
QΓh∇
δF
δh
δF
δh
+ QhΓ∇
+ QΓΓ∇
δF
δ Γ
δF
δ Γ
(6)
(7)
(cid:19)
(cid:32) h3
=
3η
h2Γ
2η
h2Γ
2η
hΓ2
η + DΓ
,
(8)
with the positive definite mobility matrix [32, 42]
(cid:90)
Figure 1. Sketch of a bacterial colony covered by an insolu-
ble surfactant with concentration Γ(x, y, t). The field h(x, y, t)
describes the height of liquid and biomass. The surfactant con-
centration on the colony is higher then on the surrounding sub-
strate, resulting in gradients in the liquid-gas surface tension and
outward-pointing Marangoni flows that promote the expansion
of the colony.
colonies of Pseudomonas aeruginosa in a one-dimensional
setting [12]. However, two-dimensional hydrodynamic sim-
ulations which focus on the shapes of spreading bacterial
colonies driven by Marangoni effects have - to the best of
our knowledge - not yet been performed. Note that besides
the surfactant-induced instability, also nutrient limitation
and chemotactic effects are a possible causes for the den-
dritic morphology of bacterial colonies (for a critical review,
see 31). In this work, we present a model for the surfactant-
driven spreading of bacterial colonies, which explicitly in-
cludes wetting effects. This allows us to study the interplay
between wettability and Marangoni fluxes and their effect
on the spreading speed and morphology. In section 2 we in-
troduce the model, a passive thin-film model with insoluble
surfactants [32] supplemented by bioactive source terms. In
section 3 we present a transversal linear stability analysis
to explore the possibility of front instabilities and perform
some full numerical simulations.
II. THIN FILM MODELLING OF
SURFACTANT-DRIVEN BIOFILM SPREADING
A bacterial colony is a complex fluid, composed of wa-
ter, bacteria, nutrients and molecules, which are secreted
extracellular polymeric substances
by the bacteria, e.g.
and surfactants.
In this work, we follow a simple two-
field modelling strategy that allows for a selective study
of the influence of wettability and Marangoni fluxes on the
spreading dynamics. We treat the bacterial colony as a
thin film of height h(x, y, t) covered by insoluble surfactant
molecules of concentration Γ(x, y, t) [see Fig. 1]. To model
the surfactant-driven spreading of the colony, we supple-
ment the hydrodynamic description with bioactive growth
processes for the film height h and the surfactant concen-
tration Γ. This approach is valid in the limit of fast osmotic
equilibration between the colony and the agar substrate.[33]
Similar models - which represent just one class in the very
rich literature concerning the mathematical modelling of
bacterial colonies (for reviews see for example 34–37) - are
used to study the influence of wettability [5, 33], quorum
sensing [38, 39] and the surfactant-driven spreading of bac-
terial colonies.[12, 18]
In the following, we first present the 'passive' part of the
hydrodynamic model for thin, surfactant-covered films be-
where η denotes the viscosity of the fluid and D is the diffu-
sivity of surfactant molecules on the interface. Performing
the variations of the free energy functional and considering
the thin film limit ζ ≈ 1 gives
δF
(9)
(10)
δh = ∂hfw − ∇ [ω∇h]
δF
δΓ = ∂Γfs .
a2 Γ. Using the common ap-
with ω = fs − Γ∂Γfs = γ − kT
proximation that the change of the surface tension with
surfactant concentration is small as compared to the refer-
ence surface tension γ (i.e. ∇ [ω∇h] ≈ γ∆h ), we obtain
(11)
∂th = ∇ · [Qhh∇[∂hfw − γ∆h] + QhΓ∇(∂Γfs)]
∂tΓ ≈ ∂t Γ = ∇ · [QΓh∇[∂hfw − γ∆h] + QΓΓ∇(∂Γfs)] .
(12)
The last term of (11) corresponds to the negative of the
Marangoni flux (cid:126)jM = − kT
2ηa2 h2∇Γ - a flux in the fluid which
is driven by concentration gradients of the surfactant.
B. Model for surfactant-driven colony spreading
In order to describe the surfactant-driven spreading of
bacterial colonies, the hydrodynamic model equations for
passive fluids (11)-(12) are extended by biological growth
and production processes. Over time, bacteria will multi-
ply by cell division and possibly secrete osmolytes. This
may result in an influx of water into the colony caused by
the difference of the osmotic pressures in the film and the
underlying moist substrate.[4] We assume that this influx is
fast as compared to the growth processes, which allows us
to write biomass production and osmotic influx as one ef-
fective growth term G(h). 1 To account for processes such
as nutrient and oxygen depletion [43, 44] which naturally
limit the colony height, we introduce a critical film height
h(cid:63), which corresponds to the maximal height that can be
sustained. It can be related to a local equilibrium of verti-
cal nutrient diffusion and consumption by bacteria.[43] We
assume a simple logistic growth law
G(h) = gh
fmod(h)
(13)
which is modified locally for very small amounts of biomass
by fmod(h) in order to prevent proliferation in the ad-
sorption layer outside of the colony and accounts for the
fact that at least one bacterium is needed to start biomass
growth. 2
1 In our previous modelling approach focussing on the osmotic effects
and wettability[5, 33] in biofilms, we pointed out that in the limiting
case of fast osmotic fluxes, a model which treats water and biomass
as two individual fields can be reduced to a simplified model with
only one variable for the film height h.
2 Here, we use fmod(h) = (1 − hu
)], but other forms
of fmod(h) with the same fixed point structure give similar results.
h )[1 − exp( hs−h
ha
(cid:19)
(cid:18)
h
h(cid:63)
1 −
3
The second bioactive process that needs to be included in
the model is the production of surfactants by the bacte-
ria. Due to the small height of the colony as compared to
its lateral extension, the surfactant quickly diffuses to the
liquid-air interface. We thus assume the production rate of
surfactant P (h, Γ) to be proportional to the biofilm height,
to decrease with increasing surfactant concentration and
to cease when the local surfactant concentration reaches a
limiting value Γmax:
P (h, Γ) = p(Γmax − Γ)Θ(Γmax − Γ) · hΘ(h − hu) .
(14)
The step-functions Θ are introduced to ensure that pro-
duction only takes place inside the colony and not in the
adsorption layer and that surfactant above the maximal
concentration Γmax is not degraded.
We include biomass growth (13) and surfactant production
(14) as additional non-conserved terms into the evolution
equations (11)-(12)
3η∇(∂hfw − γ∆h)
+ kT
a2 ∇ ·
2η∇Γ
+ G(h)
(cid:105)
(cid:105)
(cid:104) h2
(cid:104) hΓ
(cid:105)
(cid:105)
(cid:104) h3
(cid:104) h2Γ
∂th =∇ ·
∂tΓ =∇ ·
(15)
(16)
3η ∇(∂hfw − γ∆h)
+ kT
a2 ∇ ·
η ∇Γ
+ kT
a2 D∆Γ + P (h, Γ)
where we have used (4) and (8).
C. Non-dimensional form of the equations
(cid:113) γ
t = τ t
To obtain a dimensionless form of the model (15)-(16)
and thereby facilitate the analysis, we introduce the scaling
fw,s = κ fw,s
(17)
where a tilde indicates dimensionless quantities. Time, en-
ergy and vertical and horizontal length scales are
h = lh
x = Lx
y = Ly
τ = L2η
κl
κ = kT
a2
l = ha
L =
κ l ,
(18)
respectively, and will be estimated quantitatively later in
Sec. III A. Inserting the scaling, into the evolution equation
results in the dimensionless biomass growth rate g = τ g,
the dimensionless surfactant production rate p = τ pl and
the wettability parameter
W =
Aa2
akBT
h2
,
(19)
which defines the relative strength of wetting as compared
to the entropic influence of the surfactant. It is connected
to the equilibrium contact angle θe of passive stationary
droplets (without bio-active terms) by θe ∝ √W so that
larger values of W result in a less wettable substrate and
larger contact angles. If not stated otherwise, we fix the
parameters to g = 10−5, p = 10−6, h(cid:63) = 20. and D = 0.01
throughout the analysis and study the effect of the wet-
tability parameter W and the maximal surfactant concen-
tration Γmax which captures e.g. the difference between a
surfactant-producing bacterial strain and a mutant strain
deficient in surfactant production.
4
Figure 2. Panels (a) to (d) show the four different types of spreading that may occur depending on the wettability parameter W
and the maximal surfactant concentration Γmax. The resp. bottom left plots show the time dependence of the mean values of the
maximal and minimal radii of the colony while the resp. top left plots show snapshots of the h(x, y) = 0.5h(cid:63) contour line at the
times indicated by filled symbols in the bottom left plots. The resp. top and bottom right plots show the film height and surfactant
distribution, respectively, at the end of the simulation. Without surfactant, the shape of the bacterial colony is circular. At high
wettability (c), the biofilm expands over the substrate with a stable circular front. Under conditions, which do not favour wetting
(d), the spreading of the colony is arrested. If the colony produces a significant amount of surfactant Γmax, spreading is promoted
by Marangoni fluxes and the circular front becomes unstable. For a high wettability (a), the front continuously advances but is
modulated. For a higher value of W (b), the colony forms pronounced fingers which expand over the substrate while the troughs
stay arrested.
III. RESULTS
In the next section, we present an analysis of the model
which focuses on the influence of surfactants and wettabil-
ity on the spreading dynamics and morphology. We start
by performing time simulations of initially circular colonies
at different parameter values W and limiting surfactant
concentrations Γmax as illustrative examples and a graphic
description of the effects. These can already be employed
to gain a qualitative understanding of the spreading be-
haviour.
In a next step, the spreading regimes are stud-
ied for planar fronts by parameter continuation techniques
[45]. This results in a more technical description and fa-
cilitates, e.g., the analysis of the emerging front instability
by a transversal linear stability analysis. The last part of
this section contains an illustrative application of the model
and exemplarily tests counter-gradients of surfactant as a
strategy to arrest the expansion of bacterial colonies.
A.
Influence of wettability and Marangoni fluxes on
the morphology of fronts in radial geometry
In a first step, the front dynamics of model (15)-(16)
is analysed by performing two-dimensional numerical time
simulations of colony growth. These reveal the influence of
wettability and the strength of the Marangoni fluxes on
the morphology of the emerging bacterial colonies. We
5
tability and thus high contact angle (high W , Fig. 2 (d)),
the spreading of the bacterial colony is arrested (type A)
and it evolves towards a steady profile of fixed extension
and contact angle.
In both cases, the production of a significant amount of sur-
factant (high Γmax) improves the capability of the bacterial
colony to expand outwards over the substrate. It results in
a higher surfactant concentration Γ(x, y) at the centre of
the colony than on the surrounding substrate which induces
an outwards flow due to the emerging surface tension gradi-
ent. For a continuously spreading colony, these Marangoni
flows increase the spreading speed and also cause modu-
lations (type M) of the circular colony shape to develop
(low W , Fig. 2 (a)). However, eventually the growth of
these undulations slows down and the tips and troughs of
the front line translate with a similar velocity over the sub-
strate. This can be clearly seen in the time evolution of
the mean values of the maximal and minimal radii of the
colony shape.
In the case of arrested spreading, the consequences of the
surfactant production are even more drastic: In the first
place it enables a horizontal expansion of the colony. Fur-
thermore, it gives rise to the formation of pronounced fin-
gers (high W , type F in Fig. 2 (b)). At large times, the fin-
ger tips spread outwards with a constant velocity whereas
the troughs of the front line stay behind at a fixed position.
A similar distinction of two types of front instabilities, for
which the shape of the evolving front modulations becomes
stationary (M) or corresponds to continuously growing fin-
gers (F), has also been made for advancing coating films
driven by gravity or shear stress.[48] The mechanism be-
hind the pronounced fingering mode found here becomes
clear when studying the distribution of surfactant on the
colony and the surrounding substrate in more detail. Fig.
3 (a) shows a height profile of a colony with pronounced
fingers at t = 107τ.
In agreement with the experimen-
tal observation [12], we find a rim in the height profile at
the edges of the colony that is particularly pronounced at
the tips of the fingers. Due to the limiting film height h(cid:63),
the centre of the colony is relatively flat. The surfactant
concentration is shown in Fig. 3 (b) and allows one to un-
derstand the formation of the pronounced fingers. In the
troughs close to the centre of the colony, the surfactant
concentration is overall high and gradients in Γ are small.
This results in only small Marangoni fluxes which do not
suffice to overcome the arrested spreading behaviour.
In
contrast, at the tips of the fingers, gradients in Gamma
and Marangoni fluxes are strong, driving the finger tips fur-
ther outwards. If the diffusion of the surfactant is not too
high, this gradient in Γ is maintained, enabling the fingers
to continuously spread over the substrate.
To see if our model predicts a reasonable spreading speed,
we estimate the scales for time and length scales by com-
paring the numerically obtained extensions with experi-
mental measurements and plug in numbers for the remain-
ing constants in the model. The typical colony height of
15− 25 µm as measured in 12, sets the vertical length scale
to l = ha ≈ 1 µm. Together with the viscosity of η ≈ 0.1 Pa
s [12] and surface tension of water γ ≈ 70 mN/m, as well as
Figure 3. Details of the colony which spreads with pronounced
fingers for W = 0.1 and Γmax = 0.5. (a) Height profile h(x, y)
at t = 107τ reveals the rims which form at the edges of the
fingers (cp. Fig. 2 (b)). (b) Surfactant concentration (colour-
ing, dashed lines represent isolines) on a finger of the colony
(solid black line). The gradient in surface tension and thus the
Marangoni flux (cid:126)jM is strong at the tips of the fingers, driving
them further outwards. In the troughs, the surfactant concen-
tration is at an overall high level so that the Marangoni flux is
weak.
2 [1 + tanh(100x)] .
employ a finite element scheme provided by the modular
toolbox DUNE-PDELAB[46, 47]. The simulation domain
Ω = [−Lx, Lx] × [−Ly, Ly] with Lx = Ly = 5000 is dis-
cretized on a regular mesh using Nx × Ny = 512 × 512
grid points and linear ansatz and test functions. The
time-integration is performed using an implicit second or-
der Runge-Kutta scheme with adaptive time step. On the
boundaries, we apply no-flux conditions for film height and
surfactant. The initial condition is given by a small nucle-
ated bacterial colony with surfactant concentration Γmax on
the colony and 0.05 × Γmax on the surrounding substrate.
The step functions in the surfactant production term (14)
are approximated by Θ(x) ≈ 1
The time simulations of (15)-(16) reveal that - depend-
ing on the wettability and the strength of the Marangoni
fluxes in the system - four qualitatively different types of
spreading behaviour can occur. These are depicted in Fig.
2 (a) to (d) for four different choices of the parameters W
and Γmax. The respective top left plots show the contours
of the colonies at equidistant times. The resp. bottom left
plots in Fig. 2 give the time dependence of the mean values
of the maximal and minimal radii of the colony to charac-
terize its shape evolution. The resp. top and bottom right
plots show the film height and surfactant distribution pro-
files at the end of the simulation.
We first discuss the spreading behaviour of the system for
low surfactant densities (low Γmax). Consistent with the
biofilm spreading model without surfactants in 33, the sys-
tem shows a non-equilibrium transition between continu-
ously spreading and arrested colonies depending on the
wettability parameter W . For small equilibrium contact
angles and high wettability (low W , Fig. 2 (c)), the bacte-
rial colony swells vertically and horizontally until the lim-
iting film height h(cid:63) is reached. Subsequently, it spreads
horizontally over the substrate with a constant speed and
a circular (type C) colony shape. In contrast, at low wet-
a typical surfactant length scale a ≈ 3 nm, we find the lat-
eral length scale L ≈ 10µm and the time scale τ ≈ 0.03s.
With the above scales, our numerically measured dimen-
sionless expansion rate of roughly 5×10−4 in Fig. 2 (a) cor-
responds to a speed of about 10µm/min, which compares
well to the experimentally found value of 5−40µm/min.[12]
To obtain a more complete picture of the front instability,
we also study the effect of the remaining parameters g, p, h(cid:63)
and D on the colony shape. We find that a small biomass
growth rate g as well as a small maximal biofilm thickness
h(cid:63) promote the instability.
Images from the direct time
simulations can be found in appendix A1.
B. Profile and velocity of fronts in planar geometry
The time simulations of the two-dimensional system have
identified the wettability parameter W and the amount of
surfactant Γmax as two key parameters which influence the
spreading of the bacterial colony. To understand the sys-
tem in more detail, next we investigate planar fronts. In
this geometry, it is possible to perform a more systematic
analysis of the system using parameter continuation. This
technique allows for a direct observation of the influence of
W and Γmax on the front profile and velocity. To that end,
the evolution equations (15)-(16) are transformed into the
co-moving coordinate system with a constant velocity v
∂th =∇ · [Qhh∇[∂hfw − γ∆h] + QhΓ∇(∂Γfs)]
+ G(h) + v∂xh
=F1(∇, v)[h, Γ]
∂tΓ =∇ · [QΓh∇[∂hfw − γ∆h] + QΓΓ∇(∂Γfs)]
+ P (h, Γ) + v∂xΓ
(20)
(21)
=F2(∇, v)[h, Γ] ,
where we introduced F1,2(∇, v) as a short hand notation
for the nonlinear operators defined by the right-hand sides
of the evolution equations (20) and (21). In the co-moving
frame, planar fronts (h0(x), Γ0(x)) which depend only on
one spatial coordinate, x, and move with a stationary pro-
file and velocity v correspond to steady solutions
∂th0(x) = F1(∇, v)[h0(x), Γ0(x)] = 0
∂tΓ0(x) = F2(∇, v)[h0(x), Γ0(x)] = 0 ,
(22)
(23)
and can thus be analysed by parameter continuation tech-
niques [49, 50]. To that end, we use the software package
AUTO-07p[45] which has previously been successfully
employed for thin-film models, e.g.
for dewetting simple
and complex liquids [51], pattern formation in dip-coating
[52] or osmotically spreading biofilms.[5, 33] We impose
that far away from the colony, the surfactant concentration
is fixed to a small but finite value.
Fig. 4 (a) and (b) show the front profiles (h0(x), Γ0(x)) for
the parameter combinations corresponding to modulated
(M) and circular (C) spreading in the radial geometry
(Fig 2). Behind the spreading front, h and Γ reach their
saturation values h(cid:63) and Γmax, respectively. The height
profile of the front shows a typical capillary rim. The
6
Figure 4. Shape and velocity of planar fronts.
(a) and (b):
Front profiles for parameter combinations corresponding to the
circular (C) colony (W = 0.03, Γmax = 0.02) and the colony
with the modulated (M) front shape (W = 0.03, Γmax = 0.5).
The front velocity v strongly depends on the maximal surfactant
amount Γmax (c) as well as the wettability parameter W (d).
surfactant diffuses in front of the moving colony, resulting
in a linear decay. This is a typical profile, as discussed
more generally in 53 for a moving source of surfactant.
The velocity of the front is strongly affected by the surfac-
tant concentration Γmax and the wettability parameter W
(see Figs. 4 (c) and (d)). For a bacterial colony with very
small Γmax, e.g. a mutant strain deficient in surfactant
production, the front velocity is roughly a factor of two
smaller then in a colony with Γmax = 0.5. This shows
that the Marangoni effect gives a strong contribution to
the outward flux that results in the colony expansion.
In analogy to the transition from continuous to arrested
spreading observed in biofilms without Marangoni flows
[33], the biofilm expansion slows down as the conditions
do not favour wetting for large W .
In the time simulations for radial geometry discussed
in section III A, we found that the surfactant not only
influences the spreading speed of the colonies, but also
affects its morphology. We will focus on this aspect in the
next section.
C. Transversal linear stability analysis of
two-dimensional planar fronts
Now, we analyse the evolution of the morphology of pla-
nar fronts in a two-dimensional geometry. To that end, we
perform a linear transversal stability analysis employing the
ansatz
h(x, y, t) = h0(x) + h1(x) exp(ikyy + σt)
Γ(x, y, t) = Γ0(x) + Γ1(x) exp(ikyy + σt)
(24)
(25)
01000200030004000x/L05101520h0/l(a)(c)(b)(d)MC01000200030004000x/L0.00.10.20.30.40.50.6Γ00.00.10.20.30.40.5Γmax01234v×10−4W=0.03CM0.000.040.08W01234v×10−4Γmax=0.5M7
with (cid:28) 1. This corresponds to fronts consisting of
a y-invariant base state given by the stationary fronts
(h0(x), Γ0(x)) plus a small perturbation with x-dependence
(h1(x), Γ1(x)) which is modulated in the y-direction with a
wavenumber ky and grows or decays exponentially in time
with the rate σ.
Inserting this ansatz into the evolution
equations (20)-(21) one obtains to O() the linear eigen-
value problem
σh1(x) = F (cid:48)
1Γh0(x),Γ0(x)Γ1(x)
2Γh0(x),Γ0(x)Γ1(x) ,
(26)
(27)
1hh0(x),Γ0(x)h1(x) + F (cid:48)
2hh0(x),Γ0(x)h1(x) + F (cid:48)
σΓ1(x) = F (cid:48)
for eigenvalues σ and eigenfunctions (h1(x), Γ1(x)), where
F (cid:48)
i,h and F (cid:48)
i,Γ are operators denoting the Fréchet-
derivatives of the non-linear operator Fi with respect to
h and Γ, respectively. Statements about the linear stabil-
ity of the front (h0(x), Γ0(x)) can now be made determin-
ing the largest eigenvalue σ which tells if the perturbation
grows (for σ > 0) or decays (for σ < 0) in time. The linear
eigenvalue problem (26)-(27) is again solved using continu-
ation techniques. The set of equations for the steady front
profiles h0(x) and Γ0(x) employed in section III B is sup-
plemented by a set of equations for the eigenfunctions that
fulfill the same boundary conditions as the base state. This
approach, in which transversal wave number and eigenvalue
are treated as parameters in an pseudo-arclength continu-
ation, is presented in tutorial form in Ref. 54.
We again investigate the front profiles for two parameter
sets which correspond to the modulated (M) and circular
(C) spreading in the radial geometry (Fig. 2). Recall that
the respective base states (h0(x), Γ0(x)) are displayed in
Fig. 4. In order to determine the transversal stability of
these fronts, one needs to analyse the corresponding disper-
sion relations σ(ky) which are shown in Fig. 5 (a). For the
parameter set (C) with only a small concentration of surfac-
tant Γmax = 0.02, the dispersion relation decays monoton-
ically (red solid line in Fig. 5 (a)). The largest eigenvalue
is σmax = 0 at ky = 0 and the front is thus transversally
stable. The eigenfunction (h1(x), Γ1(x)) corresponding to
the largest eigenvalue (red solid lines in Fig. 5 (b),(d))
is the neutrally stable (Goldstone) mode representing the
translational symmetry of the equations. As expected, it is
identical to the spatial derivative of the front profiles (data
not shown).
For the other parameter set (M), which corresponds to the
situation that a significant amount of surfactant Γmax = 0.5
is present in the system, the largest eigenvalue is positive at
finite wavenumber (σmax = 3.98×10−7 at ky = 4.89×10−3)
and the front is thus transversally unstable (blue dashed
line in Fig. 5 (a)). These values roughly agree with linear
results extracted from the fully nonlinear time simulation
in a planar (data not shown). The eigenfunctions corre-
sponding to the largest eigenvalue (blue dashed lines in 5
(b),(d)) are strongly localized in the front region (compare
to Fig. 4). To find the surfactant concentration at which
the transition from transversally stable to unstable fronts
takes place, we follow the maximum of the dispersion rela-
tion σmax while varying Γmax (Fig. 5 (c)). At W = 0.03, we
Figure 5. Linear transversal stability analysis for planar fronts
in the co-moving frame for W = 0.03. The dispersion relation
shown in (a) monotonically decreases for Γmax = 0.02 (red) but
has a maximum σmax > 0 for Γmax = 0.5 (blue) indicating a
front that is unstable with respect to transversal perturbations.
(b) and (d) show the eigenfunctions for the fronts profiles in Fig.
4 corresponding to the transversal wave number ky with the
largest eigenvalue. The grey arrow indicates the front position
of the corresponding h profiles. Following the maximum of the
dispersion relation σmax depending on Γmax in (c) shows that
for W = 0.03, the front profile is unstable for Γmax > 0.0268.
find that σmax is positive and the front profile thus transver-
sally unstable for Γmax > 0.0268.
D. Comparison to surfactant-driven spreading of
'passive' thin films
The front profiles observed in our model show some
of the main characteristics of the solutions observed for
surfactant-driven spreading of passive thin liquid films
as e.g.
the capillary rim near the edge of the front
and the linear decay of the surfactant concentration
in front of the drop.
In contrast to other modelling
approaches [29, 30, 55], we incorporate a wetting energy
corresponding to a partially wetting fluid, resulting in a
stable adsorption layer of height ha independent of the
surfactant concentration Γ. Therefore, we do not observe
the typical fluid step or a thinned region in front of the
advancing colony which is often described as the origin
of the front instability observed for surfactant-driven
spreading of 'passive' drops on horizontal substrates.[20]
Instead, our model for surfactant-driven colony spreading
shows similarity to a surfactant covered drop sliding down
an inclined substrate.[56–58] In this set-up, the spreading
of the drop is - in addition to the Marangoni fluxes - driven
by gravity which acts as a body force on the fluid.
In
our model, the driving is presented by the non-conserved
biomass growth term. The fingering instability and the
eigenfunctions of the unstable mode observed in our model
strongly resemble the transversal perturbations found
01000200030004000x/L0.000.050.100.15h1/l(a)(b)(c)(d)MC01000200030004000x/L−1012Γ1×10−40.000.010.02ky−5−4−3−2−101σ×10−60.00.10.20.30.40.5Γmax01234σmax×10−7W=0.03in a constant-flux configuration in 57 which are also lo-
cated at the front edge rather than in the region ahead of it.
E. Phase-diagram for planar fronts
We complete our analysis of planar fronts by combin-
ing our results from the transversal linear stability analysis
with time simulations. This allows us to determine a phase-
diagram which distinguishes between different spreading
modes depending on the wettability parameter W and the
surfactant concentration Γmax. The time simulations are
performed on a domain Ω = [−Lx, Lx] × [−Ly, Ly] with
Lx = 1500 and Ly = 3000 discretized on an equidis-
tant Nx × Ny = 256 × 512 mesh with the same integra-
tion method and boundary conditions as applied in sec-
tion III A. The initial condition consists of a noisy planar
front given by the corresponding stationary front profile
(h0(x), Γ0(x)) for each parameter set. The simulation time
is Tend/τ = 107. For initially planar fronts, we find three
different spreading regimes as shown in Fig. 6: Arrested
planar fronts, which do not advance (grey dots), moving
planar fronts (red triangles) and moving modulated fronts
(blue squares) for which the transversal perturbations ∆h
grow in time (and thus ∆h(t=Tend)
∆h(t=0) > 1). We find the same
tendencies as observed in the circular geometry in section
III A. At low surfactant concentration, the front spreads
without a transversal instability for a small contact angle
(low W ) but is arrested for high W . An increased surfac-
tant concentration leads to a modulated front. We com-
pare this findings with the predictions from the transversal
linear stability analysis. We identify the region in which
the ansatz (24)-(25) of moving fronts is valid (stationary h
and Γ in the co-moving frame) is valid. In the grey region
to the right of the dashed line in Fig. 6, this condition
breaks down and we do not expect a stationary moving
front. This is in accordance with the occurrence of the ar-
rested mode in the time simulations. Note that in this sit-
uation, the produced surfactant still spreads outwards and
the arrested growth mode does therefore not correspond
to a stationary front with v = 0 for both fields h and Γ.
The transversal linear stability makes a prediction about
the strength of the transversal instability via the largest
eigenvalue σmax. To the left of the dotted line in Fig. 6,
the eigenvalue is larger than σmax = 10−8 and the modula-
tion of the front should be observable within our simulation
time Tend/τ = 5 · 106..107. This is in good agreement with
the time simulations.
Interestingly, the fingering mode (F) does not occur for
time simulations initiated with planar fronts that are only
slightly perturbed.
In general, the transversal instability
appears to be much weaker then observed in the radial ge-
ometry. This can be attributed to a dilution effect of the
surfactant: In the radial geometry, the produced surfactant
is diluted more strongly when it spreads outwards from the
colony and the surfactant profile decays faster with the dis-
tance to the front. This results in stronger gradients in
surfactant concentration which drive the transversal insta-
8
Figure 6. Phase-diagram for the spreading of two-dimensional
planar fronts. To the left (right) of the dashed line, the fronts
of the height profile h are moving (not moving). The linear
stability analysis predicts modulated moving fronts to the left
of the dotted line (the largest eigenvalue σmax is > 10−8). In
numerical time simulations initiated with a noisy planar front,
three types of spreading occur: transversally stable planar fronts
(red triangles), modulated fronts (blue squares) and arrested
fronts which can not expand over the substrate (grey dots). The
grey dots with white circles mark parameter sets for which a
finite perturbation in the initial condition leads to spreading of
a pronounced finger.
bility. To test if the fingering mode only exists for circu-
lar colonies, we perform the time simulations with an ini-
tial condition consisting of a planar front with a finite-size
perturbation in the form of a small finger. We find that
at large surfactant concentrations, the arrested spreading
mode can be overcome (grey dots with white circle in Fig.
6): the initial finger continuously grows while the rest of
the front stays behind similar to the radial geometry. In
conclusion, the instability and especially the fingering mode
are generically occurring in the planar and the radial ge-
ometry, however, the onset and strength critically depend
on colony shape.
F. Preventing the growth of bacterial colonies by a
counter-gradient of surfactant
After analysing the model mathematically in sections
III B to III E, we illustrate the consequences of the spread-
ing mechanism at an example. One strategy that has been
suggested to arrest the expansion of a bacterial colony are
counter-gradients of surfactants.
Indeed, experiments[12,
14] show that spreading of a P. aeruginosa colony can be
inhibited if exogenously added bio-surfactant is present on
the agar substrate in a circular pattern around the colony
with a concentration comparable to the in vivo one. To
test this strategy in our model, we perform a time sim-
ulation with W = 0.1 and Γmax = 0.5 starting from an
initial condition given by a surfactant-laden colony at the
centre of the simulation domain and additional surfactant
to the left of the colony. The time simulation (see Fig. 7)
shows that the growth of the colony towards the left hand
side slows down as soon as the colony 'senses' the addi-
0.000.020.040.060.080.10W0.00.10.20.30.40.5Γmaxplanararrestedmodulatedtional surfactant and eventually its growth is arrested. At
the other side, the colony performs the expected finger-like
growth. This effect can also be expected to occur when
two surfactant-producing colonies approach each other, as
observed experimentally[19].
IV. CONCLUSION AND OUTLOOK
We have developed and studied a simple model
for
surfactant-driven biofilm spreading which demonstrates
that wettability and Marangoni fluxes have a strong
effect on the expansion behaviour and morphology of
bacterial colonies. The model we have presented is based
on a hydrodynamic approach including wetting forces
supplemented by bioactive terms and thus allows us to
study the interplay between biological growth processes
and passive surface forces. We find four different types of
spreading, ranging from arrested spreading over circular
spreading and undulated spreading fronts to the formation
of pronounced fingers. The obtained results show that
the production of bio-surfactants can enable a bacterial
colony to spread over the substrate under conditions,
which are otherwise unfavourable to a horizontal expan-
sion; This is because the resulting Marangoni fluxes can
significantly contribute to the spreading velocity. Our
results are in qualitative agreement with the experimental
findings [12, 14, 19] which show that surfactant-producing
Pseusomona aeruginosa colonies spread outwards and
form pronounced fingers whereas a wild type deficient in
surfactant production can not expand and is arrested in
a small circular shape. This corresponds to the transition
from the fingering mode to the arrested spreading mode
that our model predicts.
Note that, here, capillarity and wettability on the one
hand and the gradients in surface tension on the other hand
have been treated separately in order to discuss their re-
spective effects. In an experiment with bacterial colonies
on agar plates, the wettability also depends on the concen-
tration of surfactants because they alter the surface tension
and thus the Hamaker constant (entering the parameter
W )[59]. Therefore, the difference between a bacterial strain
deficient in surfactant production and a surfactant produc-
ing strain implies that the latter has a lower parameter W
and a higher surfactant concentration Γmax (see discussion
in 33).
In this work, we have followed a simple two-field approach,
treating the bacterial colony as a complex fluid covered by
surfactants. To capture situations, in which variations of
the colony composition are not negligible, e.g. because of
similar time scales for biomass growth and osmotic pro-
9
cesses, the model can be extended to a three-field model.
There, the water concentration enters as a separate field as
described in 33. In addition, the extension of the model to
soluble surfactants with a bulk concentration is straight for-
ward, following the model for passive fluids presented in 41.
Our modelling approach neglects complex features, such as
vertical gradients or cell differentiation. Experiments [12]
Figure 7. Inhibition of bacterial expansion by a counter-gradient
of surfactant. Initially, surfactant is deposited on the left bor-
der of the computational domain, while a small drop of a bac-
terial colony is initiated in the centre. The counter-gradient of
external surfactant inhibits fingering towards the left border.
Parameters are W = 0.1 and Γmax = 0.5.
in which the rhamnolipid production in Pseudomona aerug-
inosa colonies is highlighted by autofluorescence indicate
that there are only small spatio-temporal variations in sur-
factant production throughout the colony but in general,
cell differentiation is an important phenomenon in bacte-
rial colonies and biofilms[60]. In future extensions of the
model, one may also incorporate the quorum sensing role of
the bio-surfactants which allows for a basic form of commu-
nication between individual cells. However, as our model
focuses on the physical effects of bio-surfactants, it is well
suited to show that these suffice to induce the striking fin-
gering colony shapes which are observed experimentally.
CONFLICT OF INTEREST
There are no conflicts to declare.
ACKNOWLEDGEMENT
We thank the DAAD, Studienstiftung des deutschen
Volkes, Campus France (PHC PROCOPE grant 35488SJ)
and the CNRS (grant PICS07343) for financial support. LI-
Phy is part of LabEx Tec 21 (Invest. l'Avenir, grant ANR-
11-LABX-0030).
[1] R. M. Donlan, Emerg. Infect. Dis. 8, 881 (2002).
[2] P. Stoodley, K. Sauer, D. G. Davies, and J. W. Costerton,
Annu. Rev. Microbiol. 56, 187 (2002).
−Ly0Lyy/Lt=1×106τ−Lx0Lxx/L−Ly0Lyy/Lt=5×106τ−Lx0Lxx/Lt=1×107τ01020h(x,y)−Lx0Lxx/L0.00.20.4Γ(x,y)[3] A. Yang, W. S. Tang, T. Si, and J. X. Tang, Biophys. J.
[33] S. Trinschek, K. John, S. Lecuyer, and U. Thiele, Phys.
112, 1462 (2017).
[4] A. Seminara, T. Angelini, J. Wilking, H. Vlamakis,
S. Ebrahim, R. Kolter, D. Weitz, and M. Brenner, Proc.
Natl. Acad. Sci. U. S. A. 109, 1116 (2012).
and U. Thiele, AIMS Materials
[5] S. Trinschek, K. John,
Science 3, 1138 (2016).
[6] G. E. Dilanji, M. Teplitski, and S. J. Hagen, Proc. R. Soc.
B 281, 1784 (2014).
[7] J. Yan, C. D. Nadell, H. A. Stone, N. S. Wingreen, and
B. L. Bassler, Nature Comm. 8, 327 (2017).
[8] E. Z. Ron and E. Rosenberg, Environ. Microbiol. 3, 229
(2001).
[9] J. M. Raaijmakers, I. De Bruijn, O. Nybroe, and M. On-
gena, FEMS Microbiol. Rev. 34, 1037 (2010).
[10] W.-J. Ke, Y.-H. Hsueh, Y.-C. Cheng, C.-C. Wu, and S.-T.
Liu, Front Microbiol 6, 1017 (2015).
[11] V. Leclère, R. Marti, M. Béchet, P. Fickers, and P. Jacques,
Arch. Microbiol. 186, 475 (2006).
[12] M. Fauvart, P. Phillips, D. Bachaspatimayum, N. Ver-
straeten, J. Fransaer, J. Michiels, and J. Vermant, Soft
Matter 8, 70 (2012).
[13] R. De Dier, M. Fauvart, J. Michiels, and J. Vermant, "The
role of biosurfactants in bacterial systems," in The Physi-
cal Basis of Bacterial Quorum Communication, edited by
S. Hagen (Springer, 2015) Chap. The Role of Biosurfactants
in Bacterial Systems, pp. 189–204.
[14] N. C. Caiazza, R. M. Shanks, and G. O'toole, J. Bacteriol.
187, 7351 (2005).
[15] R. Daniels, S. Reynaert, H. Hoekstra, C. Verreth,
J. Janssens, K. Braeken, M. Fauvart, S. Beullens, C. Heus-
dens, I. Lambrichts, et al., Proc. Natl. Acad. Sci. USA 103,
14965 (2006).
[16] A. Be'er, R. S. Smith, H. Zhang, E.-L. Florin, S. M. Payne,
and H. L. Swinney, J. Bacteriol. 191, 5758 (2009).
[17] R. F. Kinsinger, M. C. Shirk, and R. Fall, J. Bacteriol.
185, 5627 (2003).
[18] T. Angelini, M. Roper, R. Kolter, D. A. Weitz, and M. P.
Brenner, Proc. Natl. Acad. Sci. USA 106, 18109 (2009).
[19] J. Tremblay, A.-P. Richardson, F. LÃľpine, and E. DÃľziel,
Environ. Microbiol. 9, 2622 (2007).
[20] O. K. Matar and R. V. Craster, Soft Matter 5, 3801 (2009).
[21] A. Marmur and M. D. Lelah, Chem. Eng. Commun. 13,
[22] S. M. Troian, X. L. Wu, and S. A. Safran, Phys. Rev. Lett.
133 (1981).
62, 1496 (1989).
[23] S. He and J. Ketterson, Phys. Fluids 7, 2640 (1995).
[24] M. Cachile, A. Cazabat, S. Bardon, M. Valignat,
F. Vandenbrouck, Colloids Surf., A 159, 47 (1999).
10
Rev. Lett. 119, 078003 (2017).
[34] Q. Wang and T. Zhang, Solid State Comm. 150, 1009
(2010).
[35] I. Klapper and J. Dockery, SIAM Rev. 52, 221 (2010).
[36] H. Horn and S. Lackner, in Productive Biofilms, Advances
in Biochemical Engineering/Biotechnology, Vol. 146, edited
by K. Muffler and R. Ulber (Springer International Publish-
ing, 2014) pp. 53–76.
[37] C. Picioreanu and M. Van Loosdrecht,
in
medicine, industry and environmental biotechnology - char-
acteritics, analysis and control,"
(IWA Publishing, 2003)
Chap. Use of mathematical modelling to study biofilm de-
velopment and morphology, pp. 413–438.
"Biofilms
[38] J. Ward and J. King, J. Eng. Math. 73, 71 (2012).
[39] J. Ward, J. King, A. Koerber, P. Williams, J. Croft, and
R. Sockett, IMA J. Math. Appl. Med. Biol. 18, 263 (2001).
[40] L. M. Pismen, Patterns and interfaces in dissipative dynam-
ics (Springer Science & Business Media, 2006).
[41] U. Thiele, A. Archer, and L. Pismen, Phys. Rev. Fluids 1,
083903 (2016).
[42] M. Wilczek, W. B. H. Tewes, S. V. Gurevich, M. H. Köpf,
L. Chi, and U. Thiele, Math. Model. Nat. Phenom. 10, 44
(2015).
[43] W. Zhang, A. Seminara, M. Suaris, M. P. Brenner, D. A.
Weitz, and T. E. Angelini, New J. Phys. 16, 015028 (2014).
[44] L. Dietrich, C. Okegbe, A. Price-Whelan, H. Sakhtah,
R. Hunter, and D. Newman, J. Bacteriol. 195, 1371 (2013).
[45] E. J. Doedel and B. E. Oldeman, AUTO07p: Continuation
and Bifurcation Software for Ordinary Differential Equa-
tions, Concordia University, Montreal (2009).
[46] P. Bastian, M. Blatt, A. Dedner, C. Engwer, R. Klöfkorn,
R. Kornhuber, M. Ohlberger, and O. Sander, Computing
82, 103 (2008).
[47] P. Bastian, M. Blatt, A. Dedner, C. Engwer, R. Klöfkorn,
R. Kornhuber, M. Ohlberger, and O. Sander, Computing
82, 121 (2008).
[48] M. H. Eres, L. W. Schwartz, and R. V. Roy, Phys. Fluids
12, 1278 (2000).
[49] H. A. Dijkstra, F. W. Wubs, A. K. Cliffe, E. Doedel, I. F.
Dragomirescu, B. Eckhardt, A. Y. Gelfgat, A. Hazel, V. Lu-
carini, A. G. Salinger, E. T. Phipps, J. Sanchez-Umbria,
H. Schuttelaars, L. S. Tuckerman, and U. Thiele, Com-
mun. Comput. Phys. 15, 1 (2014).
[50] Y. A. Kuznetsov, Elements of applied bifurcation theory,
Vol. 112 (Springer Science & Business Media, 2013).
[51] U. Thiele, D. V. Todorova, and H. Lopez, Phys. Rev. Lett.
[53] H. Williams and O. Jensen, IMA J. Appl. Math. 66, 55
[54] U. Thiele, in Münsteranian Torturials on Nonlinear Sci-
ence: Continuation, edited by U. Thiele, O. Kamps, and
S. V. Gurevich (CeNoS, Münster, 2015) 1st ed.
[55] O. Jensen and S. Naire, J. Fluid Mech. 554, 5 (2006).
[56] B. Edmonstone, O. Matar, and R. Craster, J Eng Math
[57] B. Edmonstone, O. Matar, and R. Craster, Physica D:
Nonlinear Phenomena 209, 62 (2005).
[58] J. Goddard and S. Naire, J. Fluid Mech. 772, 535 (2015).
[59] U. Thiele, J. H. Snoeijer, S. Trinschek, and K. John, arXiv
preprint arXiv:1802.04042 (2018).
[60] H. Vlamakis, C. Aguilar, R. Losick, and R. Kolter, Genes
& development 22, 945 (2008).
[25] A. B. Afsar-Siddiqui, P. F. Luckham, and O. K. Matar,
J. Phys.: Condens. Matter 29, 014002 (2016).
and
111, 117801 (2013).
[52] M. Wilczek, J. Zhu, L. Chi, U. Thiele, and S. V. Gurevich,
[26] A. Afsar-Siddiqui, P. Luckham, and O. Matar, Langmuir
(2001).
Langmuir 19, 696 (2003).
20, 7575 (2004).
Lett. 65, 333 (1990).
[27] S. Troian, E. Herbolzheimer, and S. Safran, Phys. Rev.
[28] O. K. Matar and S. M. Troian, Phys. Fluids 11, 3232
[29] M. Warner, R. Craster, and O. Matar, Phys. Fluids 16,
50, 141 (2004).
(1999).
2933 (2004).
[30] R. Craster and O. Matar, Phys. Fluids 18, 032103 (2006).
[31] A. Marrocco, H. Henry, I. Holland, M. Plapp, S. Séror, and
B. Perthame, Math Model Nat Phenom 5, 148 (2010).
[32] U. Thiele, A. J. Archer, and M. Plapp, Phys. Fluids 24,
102107 (2012), note that a term was missed in the variation
of F and a correction is contained in the appendix of [? ].
V. APPENDIX
For completeness, we briefly investigate the influence of
the remaining dimensionless parameters and test the ro-
bustness of the observed phenomena by performing addi-
tional numerical time simulations. A simulation with the
parameters W = 0.05, Γmax = 0.5, g = 10−5, p = 10−6,
h(cid:63) = 20. and D = 0.01 is used as a reference and each
parameter is varied individually. Fig. 8 shows time simu-
lations which are initiated with a quarter of a small bacte-
rial colony with surfactant concentration Γmax on a domain
Ω = [0, 5000] × [0, 5000] discretized on an equidistant mesh
of Nx × Ny = 256 × 256 grid points. A larger biomass pro-
duction rate g reduces the formation of fingers as the hy-
drodynamic time-scale which is relevant for the Marangoni
fluxes that drive the instability is no longer fast enough (see
Fig.8 (a)). An increase of the limiting height h(cid:63) also results
in a weakening of the instability (see Fig.8 (b)). A change
of the surfactant diffusion D or the production rate p does
not change the morphology of the colonies drastically (see
Fig.8 (b) and (c), respectively).
11
12
Figure 8. Parameter study to test the influence of the remaining parameters g, h(cid:63), D and p on the colony shape. The contour
line h(x, y) = 0.5h(cid:63) is shown at equidistant times with ∆t = 106. In each run, one parameter is varied as compared to a reference
parameter set (middle column) with W = 0.05, Γmax = 0.5, g = 10−5, p = 10−6, h(cid:63) = 20. and D = 0.01.
0Lx0Lyy/L(a)(a)(a)(a)(a)g=5×10−60Lx0Lyg=1×10−50Lx0Lyg=2×10−50Lx0Lyy/L(b)(b)(b)(b)(b)h?=100Lx0Lyh?=200Lx0Lyh?=400Lx0Lyy/L(c)(c)(c)(c)(c)D=0.0010Lx0LyD=0.010Lx0LyD=0.10Lxx/L0Lyy/L(d)(d)(d)(d)(d)p=5×10−70Lxx/L0Lyp=1×10−60Lxx/L0Lyp=2×10−6 |
1201.4300 | 1 | 1201 | 2012-01-20T14:20:20 | Non-destructive imaging of an individual protein | [
"physics.bio-ph"
] | The mode of action of proteins is to a large extent given by their ability to adopt different conformations. This is why imaging single biomolecules at atomic resolution is one of the ultimate goals of biophysics and structural biology. The existing protein database has emerged from X-ray crystallography, NMR or cryo-TEM investigations. However, these tools all require averaging over a large number of proteins and thus over different conformations. This of course results in the loss of structural information. Likewise it has been shown that even the emergent X-FEL technique will not get away without averaging over a large quantity of molecules. Here we report the first recordings of a protein at sub-nanometer resolution obtained from one individual ferritin by means of low-energy electron holography. One single protein could be imaged for an extended period of time without any sign of radiation damage. Since ferritin exhibits an iron core, the holographic reconstructions could also be cross-validated against TEM images of the very same molecule by imaging the iron cluster inside the molecule while the protein shell is decomposed. | physics.bio-ph | physics | Non-destructive imaging of an individual protein
J.-N. Longchamp*, T. Latychevskaia, C. Escher, H.-W . Fink
Physics Institute
University of Zurich
W interthurerstrasse 190
8057 Zurich
Switzerland
*Corresponding author:
E-mail: [email protected]
Major category: Physics
Minor category: Biophysics and Computational Biology
Abstract
The mode of action of proteins is to a large extent given by their ability to adopt
different conformations. This is why imaging single biomolecules at atomic resolution
is one of the ultimate goals of biophysics and structural biology. The existing protein
database has emerged from X-ray crystallography, NMR or cryo-TEM investigations.
However, these tools all require averaging over a large number of proteins and thus
over different conformations. This of course results in the loss of structural
information. Likewise it has been shown that even the emergent X-FEL technique will
not get away without averaging over a large quantity of molecules(1).
Here we report the first recordings of a protein at sub-nanometer resolution obtained
from one individual ferritin by means of low-energy electron holography. One single
protein could be imaged for an extended period of time without any sign of radiation
damage(2). Since ferritin exhibits an iron core, the holographic reconstructions could
also be cross-validated against TEM images of the very same molecule by imaging
the iron cluster inside the molecule while the protein shell is decomposed .
Introduction
All conventional methods for investigating the structure of biomolecules such as
proteins suffer from significant drawbacks. X-ray crystallography, NMR and cryo-
electron microscopy require averaging over a large number of molecules, usually
more than 106 similar entities(6) and as a consequence conformational details remain
undiscovered. Furthermore, these methods can only be applied to a small subset of
biological molecules that either readily crystallize to be used for X-ray studies or are
readily soluble and small enough (< 35 kDa) for NMR investigations.
Emerging X-ray coherent diffraction imaging (CDI) projects envision imaging the
molecule instantaneously before it is destroyed by radiation damage(7). Recently,
Seibert et al. succeeded in imaging a single highly symmetric 750 nm diameter
mimivirus with 32 nm resolution(8). While there was the initial believe that this
technique would provide high resolution images of a single biomolecule, it is now
evident that averaging over a large number of molecules cannot be omitted even with
the brightest source of the X-FEL projects(1, 9).
We have recently demonstrated that electrons with kinetic energies in the range of
50-200 eV do not cause radiation damage to biomolecules, enabling the investigation
of an individual molecule for an extended period of time(2). Since the electron
wavelength associated with this kinetic energy range is between 0.86 Å (for 200 eV)
and 1.7 Å (for 50 eV), low-energy electrons have the potential for non-destructive
imaging of single biomolecules and in particular individual proteins at atom ic
resolution.
Our low-energy electron holography experimental scheme transcribes the original
idea of holography invented by Dennis Gabor(10). A sharp tungsten tip acts as field
emitter of a divergent coherent electron beam (11). As illustrated in Figure 1a, the
sample is brought in the path of the coherent electron wave. The electrons scattered
off the object interfere with the unscattered so-called reference wave and at a distant
detector the interference pattern is recorded(12). A hologram, in contrast to a
diffraction pattern, contains the phase information of the object wave and the object
can be unambiguously reconstructed(13).
The protein of interest needs to be free-standing in space when exposed to the low-
energy electron beam. Therefore the protein is attached to a carbon nanotube
suspended over a hole. Strictly speaking, arrays of small holes of 240 nm in diameter
are milled in a carbon coated silicon nitride membrane by means of a focused gallium
ion beam. The nanotube-ferritin complex is kept in aqueous solution of which a
droplet is applied onto the holey membrane. Once the droplet has dried out some of
the nanotubes remain across holes providing free-standing ferritin molecules. These
molecules can then be examined in our low-energy electron microscope(12).
Prior to the described preparation procedure the carbon nanotubes undergo acid
treatment in order to form carboxyl groups on the outer wall and hence disperse
efficiently in ultra highly purified water. Adding a buffered solution of proteins, the
latter eventually bind to the nanotubes by dipole forces, as schematically illustrated in
the inset of Figure 1. This preparation does not rely on specific features of ferritin and
hence is applicable to a large class of biomolecules.
Here we report the first images of ferritin at sub-nanometer resolution obtained from
one individual protein by means of low-energy electron holography. The proteins
could be imaged for an extended period of time without any sign of radiation damage.
Ferritin is a globular protein with a molecular weight of 450 kDa and is composed of
24 subunits, the outer diameter of the protein amounts to 11 -12 nm and the inner
cavity exhibits a diameter of 8 nm(3-4). Up to 4500 iron atoms can be stored in the
cavity and the main function of ferritin in the animal metabolism is the regulation of
the iron level in the body(4). Recently, it has been shown that ferritin plays an
essential role in the progression of neuronal degenerative diseases like Alzheimer or
Parkinson(5). Understanding the function of ferritin in more detail by identifying its
various conformations, would be an important step towards a better comprehension
of the mechanism behind these diseases.
Results
In Figure 2a, an example of a low-energy electron hologram of ferritin molecules
attached
to nanotubes
is presented. There are several
ferritin molecules
chemisorbed onto the nanotubes as evident from the hologram reconstruction.
Individual ferritin molecules can be resolved, as shown in Figure 2b-c. The ability to
image several molecules exhibiting different orientations and conformations in just
one single-shot is a promising feature inherent to our method. The individual ferritin
molecules have been monitored for more than 20 minutes without observing any
change in the hologram. The corresponding total dose amounts to more than 10 9
electrons/nm2 and is at least six orders of magnitude higher than the permissible
dose in TEM or X-ray examinations of biomolecules. The globular form and the size
are in agreement with the structure proposed by X-ray crystallography and cryo-TEM
investigations.
Due to the large amount of iron nuclei stored in the center of the molecule, ferritin can
also be detected by high-energy electron imaging techniques(14-17) without any
additional heavy metal staining. Though these techniques lead to radiation damage
of the protein shell they allow for subsequent cross-validation in terms of presence
and location of the molecule.
In Figure 3a, a hologram of a ferritin sample recorded with electrons of 57 eV kinetic
energy is depicted. The presence of a bundle of carbon nanotubes as well as the
globular structures on its outer wall can already be identified before the numerical
reconstruction has been performed. The result of the latter is presented in Figure 3b.
The right side of the nanotube corresponds to the correct representation, free of twin-
image artifacts(18), as obtained by single-sideband holography reconstruction(19). A
close-up of the reconstruction of an individual ferritin molecule is displayed in Figure
3c.
For cross-validation, the sample was subsequently examined in a conventional 80 kV
TEM. Figure 3c displays the TEM image of the very same sample as shown in Figure
3a-c. The 8 nm iron core of the ferritin is apparent at the very same position as in the
previous holographic reconstruction. While the position of the ferritin observed in the
low-energy point source microscope perfectly corresponds to that observed in the
TEM, the way the protein is attached to the nanotube appears to be different. In
contrast to the holographic reconstruction where the protein appears as a sphere
attached to an elongated structure, the TEM image displays the protein as if it was
“melded” with the nanotube. The overall size of the protein observed in the TEM
amounts to 9-10 nm. We associate theses differences to the radiation damage which
has occurred during TEM imaging. Following the high-energy electron investigations,
the sample was transferred back to our low-energy electron point source microscope;
the corresponding hologram and its reconstruction are displayed in Figure 3e-f
respectively. Here, the damage provoked by the 80 keV kinetic energy electrons is
evident. The way the protein is now attached to the nanotubes coincides with the
TEM results.
Discussion
We have reported the very first non-destructive investigation of an individual protein
by means of low-energy electron holography. The successful imaging of a single
protein is cross-validated by TEM examination. The sample preparation method can
be applied to a broad class of molecules.
While the spatial resolution in the holographic reconstructions presented here is
already below the nanometer scale, it is now an experimental challenge to reach a
resolution close to the employed wavelength with an advanced low-energy electron
holographic setup.
Besides
this engineering challenge of
improving
interference
resolution
in
holography, another route towards higher resolution is to combine holography with
coherent diffraction imaging using low-energy electrons(20). Essential ingredients like
a micron sized electron lens(21) and the ability to record coherent diffraction
patterns(22) are already at hand for the pursuit of single molecule structural biology
at atomic resolution.
Methods
Sample preparation. COOH-functionalized double-walled carbon nanotubes from
the company NanoLab were dispersed to a concentration of 0.2 mg/ml by sonication
for 15 min in ultra highly purified water. Subsequently, ferritin proteins in a MES
buffer (pH 6.5) were added to the carbon nanotubes solution, with a final
concentration of 0.5 nM. The ferritin-nanotube solution was kept under gentle stirring
overnight.
SiN membranes (50 nm thickness) from the company Silson were first coated with a
5 nm thick carbon layer. Subsequently, holes of 200 nm in diameter were milled in
the membranes by means of a focused Ga ion beam. Thereafter, the membranes
were coated again with 20 nm carbon layers on both sides.
Next, the membranes were hydrophylized by exposition to ozone in an UV-ozone
oven for 20 min. Before 1 l of the ferritin-nanotubes solution was applied and air-
dryed on the membrane.
Holography. The detector system of our
low-energy electron point source
microscope consists of a multi-channel plate (MCP) and a phosphor screen directly
mounted on a fiber optic plate (FOP). Both, MCP and Phosphor screen-FOP are 75
mm in diameter. A 8000x6000 pixels CCD camera was used to record the
holograms. The overall resolution of the detector system is around 20m.
An electro chemically etched single-crystal <111>-tungsten tip was used as point
source of coherent low-energy electrons. The holograms were recorded with a total
electron current of 50 nA.
Acknowledgements
We would like to thank Jennifer Clark for her help with sample preparation and the
Swiss National Science Foundation for its financial support.
Author Contributions
J.N.L. conducted the holographic and TEM investigations. T.L. performed the
hologram reconstructions. J.N.L. and C.E. worked on the sample preparation and
H.W .F. contributed to the design of the experimental setup. All authors discussed the
results and implications and jointly wrote the manuscript.
References
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Neutze, R., Wouts, R., van der Spoel, D., Weckert, E. & Hajdu, J. Potential for biomolecular
imaging with femtosecond X-ray pulses. Nature 406, 752-757 (2000).
Germann, M., Latychevskaia, T., Escher, C. & Fink, H.-W. Nondestructive imaging of individual
biomolecules. Phys. Rev. Lett. 104, 095501 (2010).
Lawson, D. M. et al. Solving the structure of human H ferritin by genetically engineering
intermolecular crystal contacts. Nature 349, 541-544 (1991).
Theil, E. C. Ferritin: Structure, Gene Regulation, and Cellular Function in Animals, Plants, and
Microorganisms. Annual Review of Biochemistry 56, 289-315,
doi:doi:10.1146/annurev.bi.56.070187.001445 (1987).
Jellinger, K., Paulus, W., Grundke-Iqbal, I., Riederer, P. & Youdim, M. Brain iron and ferritin in
Parkinson's and Alzheimer's diseases. Journal of Neural Transmission: Parkinson's Disease
and Dementia Section 2, 327-340, doi:10.1007/bf02252926 (1990).
Chiu, W. et al. High-resolution cryoelectron microscopy of biological specimens.
Ultramicroscopy 23, 232-232 (1987).
Chapman, H. N. et al. Ultrafast coherent diffraction imaging with x-ray free-electron lasers.
Proceedings of FEL 2006, BESSY, Berlin, Germany, 805-811 (2006).
Seibert, M. M. et al. Single mimivirus particles intercepted and imaged with an X-ray laser.
Nature 470, 78-81, doi:10.1038/nature09748 (2011).
Miao, J. W., Chapman, H. N., Kirz, J., Sayre, D. & Hodgson, K. O. Taking X-ray diffraction to the
limit: Macromolecular structures from femtosecond X-ray pulses and diffraction microscopy
of cells with synchrotron radiation. Annu. Rev. Biophys. Biomol. Struct. 33, 157-176 (2004).
Fink, H. W. in Electron Microscopy 1994, Vol 1 - Interdisciplinary Developments and Tools
eds B. Jouffrey & C. Colliex) 319-320 (1994).
Fink, H. W. Point-source for ions and electrons. Physica Scripta 38, 260-263 (1988).
Fink, H.-W., Stocker, W. & Schmid, H. Holography with low-energy electrons. Phys. Rev. Lett.
65, 1204-1206 (1990).
Latychevskaia, T. & Fink, H.-W. Simultaneous reconstruction of phase and amplitude contrast
from a single holographic record. Opt. Express 17, 10697-10705 (2009).
Mann, S., Bannister, J. V. & Williams, R. J. P. Structure and composition of ferritin cores
isolated from human spleen, limpet (Patella vulgata) hemolymph and bacterial
(Pseudomonas aeruginosa) cells. Journal of Molecular Biology 188, 225-232 (1986).
Kawasaki, T., Endo, J., Matsuda, T., Osakabe, N. & Tonomura, A. Applications of Holographic
Interference Electron Microscopy to the Observation of Biological Specimens. Journal of
electron microscopy 35, 211-214 (1986).
Lichte, H. & Lehmann, M. in Advances in imaging and electron physics Vol. Volume 123 eds
Pier Georgio Merli Gianluca Calestani Peter W. Hawkes & Vittori-Antisari Marco) 225-255
(Elsevier, 2002).
Quintana, C., Cowley, J. M. & Marhic, C. Electron nanodiffraction and high-resolution
electron microscopy studies of the structure and composition of physiological and
pathological ferritin. J. Struct. Biol. 147, 166-178 (2004).
Latychevskaia, T. & Fink, H.-W. Solution to the twin image problem in holography. Phys. Rev.
Lett. 98, 233901 (2007).
Bryngdah.O & Lohmann, A. Single-sideband holography. J. Opt. Soc. Am. 58, 620-624 (1968).
Latychevskaia, T., Longchamp, J. N. & Fink, H.-W. When Holography Meets Coherent
Diffraction Imaging arXiv, 1106.1320v1101 (2011).
Steinwand, E., Longchamp, J. N. & Fink, H. W. Fabrication and characterization of low
aberration micrometer-sized electron lenses. Ultramicroscopy 110, 1148-1153 (2010).
Steinwand, E., Longchamp, J.-N. & Fink, H.-W. Coherent low-energy electron diffraction on
individual nanometer sized objects. Ultramicroscopy 111, 282-284 (2011).
16
17
18
19
20
21
22
Figure legends
Figure 1: Scheme for recording the low-energy electron hologram of a protein.
Conduction electrons confined in a pointed W (111) single crystal wire are field
emitted into vacuum at an atomic-sized emission area providing a coherent low-
energy electron point source (PS). At the less than 1 micron distant object-plane
(OP), part of the coherent electron wave is scattered by a ferritin attached to a carbon
nanotube constituting the object wave indicated in red. At a distant detector screen
(S), the far-field interference pattern between object- (red) and reference-wave (blue)
- the hologram - is recorded and its digital record is subject to the numerical
reconstruction of the protein.
Figure 2: Low-energy electron hologram of ferritin and
its numerical
reconstruction. a 53eV kinetic energy electron hologram of several ferritin
molecules attached to a bundle of carbon nanotubes. b reconstruction of a. c close-
up of b.
Figure 3: Control experiments using a TEM. a 57eV kinetic energy electron
hologram of individual ferritin molecules attached to a bundle of carbon nanotubes. b
side-band holography reconstruction of a. c close-up of b. d TEM image of the very
same ferritin molecule. e 45eV kinetic energy electron hologram of the very same
sample as
in a but after
the TEM
investigations. f side-band holography
reconstruction of e.
|
1705.07821 | 4 | 1705 | 2018-05-03T13:40:57 | Model of collective fish behavior with hydrodynamic interactions | [
"physics.bio-ph"
] | Fish schooling is often modeled with self-propelled particles subject to phenomenological behavioral rules. Although fish are known to sense and exploit flow features, these models usually neglect hydrodynamics. Here, we propose a novel model that couples behavioral rules with far-field hydrodynamic interactions. We show that (1) a new "collective turning" phase emerges; (2) on average individuals swim faster thanks to the fluid; (3) the flow enhances behavioral noise. The results of this model suggest that hydrodynamic effects should be considered to fully understand the collective dynamics of fish. | physics.bio-ph | physics |
Model of collective fish behavior with hydrodynamic interactions
Audrey Filella,1 Fran¸cois Nadal,2 Cl´ement Sire,3 Eva Kanso,4 and Christophe Eloy1, ∗
1Aix Marseille Univ, CNRS, Centrale Marseille, IRPHE, 13013 Marseille, France
2Department of Mechanical, Electrical and Manufacturing Engineering,
Loughborough University, Loughborough LE11 3TU, UK
3CNRS, Universit´e Paul Sabatier, Laboratoire de Physique Th´eorique, 31062 Toulouse, France
4Aerospace and Mechanical Engineering, University of Southern California,
854 Downey Way, Los Angeles, CA 90089, USA
(Dated: May 4, 2018)
Fish schooling is often modeled with self-propelled particles subject to phenomenological behav-
ioral rules. Although fish are known to sense and exploit flow features, these models usually neglect
hydrodynamics. Here, we propose a novel model that couples behavioral rules with far-field hydro-
dynamic interactions. We show that (1) a new "collective turning" phase emerges; (2) on average
individuals swim faster thanks to the fluid; (3) the flow enhances behavioral noise. The results of
this model suggest that hydrodynamic effects should be considered to fully understand the collective
dynamics of fish.
Collective animal motion is ubiquitous: insects swarm,
birds flock, hoofed vertebrates herd, and even humans
exhibit coordination in crowds [1–5]. Among these fas-
cinating collective behaviors, schooling refers to the co-
ordinated motion of fish. It is exhibited by half of the
known fish species during some phase of their life cy-
cle [6] and can generate different disordered phases, such
as swarming (important cohesion but low polarization of
fish heading), or ordered phases: milling (torus or vor-
tex pattern), bait ball (dense "ball" of fish), or highly
polarized schools [7].
Interestingly, these collective phases can be achieved
without any leader in the group. This has first been ob-
served in experiments [8], and later confirmed by the de-
velopment of self-propelled particle (SPP) models [9, 10].
In the context of fish schooling, these mathematical mod-
els have generally been constructed by assuming simple
phenomenological behavioral rules, such as the popular
"three-A rules" of avoidance, alignment, and attraction
[11–13]. From a physical point of view, these SPP models
have an obvious interest because of their simplicity and
universality [14], and because they allow the derivation
of continuum equations [15]. Similar approaches have
been used in soft active matter (e.g., bacteria swarms
and microtubule bundles) to derive continuous models
taking into account hydrodynamic interactions at vanish-
ing Reynolds number [16, 17]. However, they have rarely
been connected quantitatively to experimental observa-
tions [18]. It is only recently that it has been possible
to infer and model the actual behavioral rules from the
individual tracking of fish in a tank [19–21].
Schooling likely serves multiple purposes [22], includ-
ing better foraging for patchy resources and increased
protection against predators. Fish are also thought to
benefit from the hydrodynamic interactions with their
neighbors [23–26], but it is unclear whether this requires
particular configurations or regulations. When swim-
ming in a structured flow, fish can exploit near-field vor-
tices generated by other fish to reduce the energetic costs
of locomotion [27–29]. Fish use their lateral line, a hair-
based sensor running along their side [30], to sense the
surrounding flow, which has also proved crucial for collec-
tive behavior [31, 32]. Yet, the existing behavioral models
of fish schooling do not include hydrodynamics. Here, we
propose to combine a data-driven attraction-alignment
model [19–21], with far-field hydrodynamic interactions.
In the context of this new model, several questions arise:
Can the swimmers exploit the flow to swim faster on av-
erage? Do the hydrodynamic interactions give rise to
novel collective phases? Does the flow play the role of
a self-induced noise, as it is the case at low Reynolds
number [33]?
Fish are modeled as self-propelled particles moving in
an unbounded two-dimensional plane. They move at con-
stant speed v relative to the flow and exhibit no inertia
[34]. Following behavioral rules inferred from shallow-
water tracking experiments [19, 20], we consider that
each individual is attracted to its Voronoi neighbors with
intensity kp (units m−1 s−1), tends to align with the same
neighbors with intensity kv (units m−1), and is subject
to a rotational noise with standard deviation σ (units
FIG. 1. (a) Sketch of two interacting swimmers, showing the
(cid:107)
heading e
i of swimmer i, its viewing angle θij, the relative
alignment angle φij, the inter-swimmer distance ρij, and the
polar coordinates in the reference frame of swimmer j (eρ
j ).
(b) Grayscale representation of the anisotropic visual percep-
tion, modeled by the term (1 + cos θij) in Eqs. (2–4).
j ,eθ
(a)(b)rad s−1/2). Moreover, each swimmer responds to the
far-field flow disturbance created by all other swimmers.
This flow is an elementary dipole, with dipole intensity
Sv in two dimensions, where S = πr2
0 is the swimmer
surface and r0 its typical length [25, 35–37]. Note that,
except for hydrodynamic interactions involving the swim-
mer size r0, swimmers are considered as point-like parti-
cles.
less, yielding the length scale (cid:112)v/kp and time scale
(cid:112)v/kp, In = σ(vkp)−1/4
1/(cid:112)vkp. To this end, I(cid:107) = kv
We use v and kp to make the problem dimension-
and If = Skp/v characterize the alignment, noise and
dipole intensities, respectively. The dimensionless equa-
tions of motion are
ri = e(cid:107)i + Ui,
θi = (cid:104)ρij sin(θij) + I(cid:107) sin(φij)(cid:105) + Inη + Ωi.
(2)
(1)
Equation (1) expresses that each individual, located at
ri, is moving with a constant unit speed along its ori-
entation e(cid:107)i (Fig. 1a). An additional drift term, Ui,
arises when hydrodynamic interactions are taken into ac-
count. A far-field approximation is used to model the
flow [37, 38]. Here, we choose to neglect the vorticity
shed in the swimmer wakes [39, 40] to keep the model
simple and tractable. Under this potential flow approxi-
mation, each swimmer generates a dipolar flow field, and
we can use the principle of superposition to calculate the
flow Ui experienced by a swimmer
Ui =
uji, with uji =
If
π
j sin θji + eρ
eθ
j cos θji
ρ2
ij
, (3)
(cid:88)
j(cid:54)=i
j ,eθ
where uji is the velocity induced by swimmer j at the
position ri and (eρ
j ) are the polar coordinates in the
framework of swimmer j (Fig. 1a). The angular velocity
in Eq. (2) is the sum of an attraction term, an align-
ment term, a standard Wiener process η(t), describing
the spontaneous motion of the fish and modeling its "free
will", and a rotational term Ωi induced by hydrodynamic
interactions. The behavioral terms (attraction and align-
ment) are averaged over the Voronoi neighbors, noted Vi,
with the weight (1 + cos θij) modeling continuously a rear
blind angle [20] (Fig. 1b)
(cid:88)
j∈Vi
(cid:30)(cid:88)
j∈Vi
(cid:104)◦(cid:105) =
◦ (1 + cos θij)
(1 + cos θij).
(4)
The rotation induced by hydrodynamic interactions is
not due to vorticity, since it is zero for a potential flow,
but to gradients of normal velocities along the swimming
direction. In other words, the angular velocity due to the
flow is
Ωi =
e(cid:107)i · ∇uji · e⊥i .
(5)
(cid:88)
j(cid:54)=i
2
FIG. 2.
(a–d) Plots of the swimmer positions and associ-
ated streamlines for different values of the parameters. On all
figures, the dipole intensity is If = 10−2, the scale bar corre-
sponds to 10 r0, and color scale represents the instantaneous
velocity. For clarity, swimmers are represented as "airfoils" of
length 7r0. Four distinct dynamical phases are observed: (a)
swarming for In = 0.8, I(cid:107) = 0.5; (b) schooling for In = 0.5,
I(cid:107) = 9; (c) milling for In = 0.3, I(cid:107) = 1.5; and (d) turning for
In = 0.2, I(cid:107) = 4. (e) Paths followed by each swimmer during
12 dimensionless time units, the final time corresponding to
(d). For long-time dynamics, see Supplementary Fig. 1 and
Supplementary Movies 1–4.
Another interpretation of this angular velocity is to
consider that each swimmer is a dumbbell oriented in the
swimming direction, with each weight advected by the
flow (Supplementary Fig. 2). The intensity of the hydro-
dynamic interactions (related to both the drift term Ui
and the induced rotation Ωi) is proportional to the dipole
intensity If . For the 10 cm-long fish (r0 = 5 cm) consid-
[19], the cruise speed is v ≈ 0.2 m s−1, and
ered in Ref.
kp = 0.41 m−1 s−1, yielding a dipole intensity If ≈ 0.016.
3
FIG. 3. Phase diagram using the value of the polarization P and the milling M as a color code (d), for three cases: (a) no
fluid (Ui = Ωi = 0); (b) fluid without induced rotation (Ui (cid:54)= 0, Ωi = 0); (c) full hydrodynamic model (Ui (cid:54)= 0, Ωi (cid:54)= 0). In all
cases, the dipole intensity is If = 10−2, the solid white line indicates the M = 0.4 level, and the dashed line the P = 0.5 level.
Solid black lines in (b–c) show the V -levels and the solid blue line in (a) shows the milling-schooling transition line found in
[20]. In (c), the black dots show the parameter values corresponding to Fig. 2a–d.
In the present model, the flow induces translational and
rotational motions, whose origin are physical, but,
in
principle, it could also elicit a behavioral response (e.g.,
a tendency for fish to go along or against the flow) [21].
We consider a group of N = 100 individuals, with
random initial orientations and initially distributed in a
20× 20 box (in dimensionless length units), although the
subsequent dynamics is not affected by the initial con-
ditions. The dynamical system described by Eqs. (1–5)
is solved numerically, using an explicit scheme with time
step δt = 10−2. Depending on the values of the three
dimensionless parameters (I(cid:107), In, and If ), four different
dynamical phases emerge (Fig. 2). When noise is com-
parable or larger than the alignment, we observe a disor-
dered swarming phase (Fig. 2a): swimmers form a sparse
group with no preferential orientation. When alignment
intensity is stronger, the group is denser and individu-
als tend to swim in the same direction: this is known
as the schooling phase (Fig. 2b). When alignment and
attraction are comparable and noise is low or moderate,
the group reaches a milling phase (Fig. 2c):
it forms a
"vortex". These three phases (swarming, schooling, and
milling) can also be observed without any hydrodynamic
interactions [20] (Ui = 0 and Ωi = 0, in Eqs. (1–2)).
However, when the flow is explicitly taken into account,
a novel phase appears that we call the turning phase
(Fig. 2d–e). In this new phase, swimmers tend to align
along a preferential orientation and, at the same time,
the group follows a large-scale quasi circular trajectory.
In order to precisely characterize these different phases,
the global order parameters P and M are introduced [41],
along with the average speed V
P = e(cid:107)i , M =
where er
i = (ri − ri)/ri − ri is the unit vector along
the segment joining the center of mass of the group and
the i-th swimmer, and the over bar denotes average over
V = ri,
(6)
(cid:12)(cid:12)er
(cid:12)(cid:12)er
i
(cid:12)(cid:12)
(cid:12)(cid:12) ,
(cid:12)(cid:12)(cid:12)(cid:12) ri
i × ri
all individuals. The parameter P is the polarization, M
is the milling and corresponds to the normalized angular
momentum of the group (straight-line schooling gives a
value of 0, while perfect milling gives 1).
To assess the importance of hydrodynamic interac-
tions, we performed a systematic parametric study for
three cases (Fig. 3): a pure behavioral model with no
effects of the fluid (Ui = 0 and Ωi = 0, in Eqs. (1–
2)), a simple model of hydrodynamic drift without in-
duced rotation (Ui (cid:54)= 0 and Ωi = 0), and a full hydrody-
namic model with both induced translation and rotation
(Ui (cid:54)= 0 and Ωi (cid:54)= 0). For each set of parameters, P , M ,
and V are obtained after time-averaging over ∆t = 100
(after waiting 100 time units to ensure that the transient
dynamics of few dimensionless time units is over), and
ensemble-averaging over 100 realizations.
In the absence of hydrodynamic interactions, the re-
sults of Ref.
[20] are recovered (Fig. 3a). For P > 0.5,
(cid:38) 2, we observe the
which roughly correspond to I(cid:107)
(cid:46) 2 and
schooling phase. For M > 0.4, obtained for I(cid:107)
In (cid:46) 0.5, the group exhibits a milling phase. For all other
cases tested, the swarming phase is observed. We chose
the threshold values P = 0.5 and M = 0.4 to distin-
guish the four phases, but P and M vary continuously in
the parameter space. An alternative choice of thresholds
would be possible and would yield qualitatively similar
results.
When hydrodynamic drift is introduced with If =
10−2, but induced rotation is neglected (Fig. 3b), the
phase diagram is practically unchanged. The only differ-
ence is that the mean velocity V is now slightly greater
than 1 (V = 1 when hydrodynamic interactions are ne-
glected). When the full hydrodynamic model is con-
sidered (Fig. 3c), the new turning phase appears for
M > 0.4 and P > 0.5 (corresponding to In (cid:46) 0.25 and
(cid:46) 5) and V is increased. Looking at the swim-
3 (cid:46) I(cid:107)
ming speeds of each individual in the group (Fig. 2e),
we see that some individuals swim slower because of the
noiseInalignmentIkpolarizationPmillingM00.5100.51noiseInnoiseIn(a)(b)(c)(d)No fluidFluid ( )Fluid ( )Color code1.2swarmingmillingschooling⌦i=0⌦i6=0Fig3turning4
and swimmers thus tend to spend more time in-line.
The role of the fluid is not only to increase the swim-
ming speed on average, but also to introduce a source
of disorder. To assess if this disorder has the same ef-
fective impact as the noise Inη in Eq. (2) that describes
the spontaneous angle fluctuations, we performed simu-
lations with no noise (In = 0) and with varying dipole in-
tensity If (Fig. 5a). The phase diagrams shown in Fig. 3c
and Fig. 5a are qualitatively similar, both exhibiting the
four phases (schooling, swarming, milling, and turning)
with the same topology. It thus shows that the hydrody-
namic interactions also play the role of a rotational noise.
There are however some differences between Fig. 3c and
Fig. 5a. First, the average velocity increases when the
dipole intensity increases whereas it tends to decrease
with noise intensity. Second, for large If and small I(cid:107),
even if P and M , are both small, the school does not
behave as in the swarming phase. It can be composed of
very dense clusters of quasi static swimmers (Supplemen-
tary Fig. 5). This non-realistic behavior is an artifact of
the simulations due to the absence of noise.
There is a continuous transition between the milling,
turning, and schooling phases (Figs. 3 and 5). In a pure
behavioral model (with no fluid), the transition between
the milling and schooling is also continuous, but M is
systematically higher when the fluid is present (Fig. 5b).
Although the turning and milling phases are similar,
their origins are different. As Calovi et al.
[20] al-
ready noted, the milling phase can only be stabilized
when swimmers have an anisotropic visual perception
(Fig. 1b and Eq. (4)). On the contrary, the turning
phase still exists when visual perception is isotropic (Sup-
plementary Fig. 7), but requires the full hydrodynamic
model to be observed. Although experimental data are
too scarce to support the existence of the turning phase
in real fish schools, we can speculate that the speed en-
hancement achieved in this phase could be advantageous
FIG. 5. (a) Phase diagram in the absence of noise (In = 0).
The color code and the contour levels are the same as in Fig. 3.
(b) Values of the polarization P (red), milling M (green),
and mean velocity V (black) for two cases: no fluid (If = 0)
and low noise (In = 0.05) with solid lines; full hydrodynamic
model (If = 10−2) and no noise (In = 0) with open symbols.
FIG. 4. Heat maps showing the probability of presence p(ρ, θ)
of all other swimmers in the framework of an individual. The
parameters are the same as in Fig. 2.
fluid, and some others swim much faster with swimming
speed reaching ri = 2.5.
To understand why individuals always swim faster on
average when hydrodynamic interactions are taken into
account, we computed the probability of presence of other
swimmers in the framework of each individual (Fig. 4).
For the same parameter values as in Fig. 2, we collected
the position and orientation of each swimmer during
∆t = 900. For the swarming phase (Fig. 4a), the proba-
bility of presence is isotropic. Hence, there is no velocity
increase due to dipolar hydrodynamic interactions on av-
erage. However, for the schooling phase, individuals tend
to swim in-line rather than side-by-side, leading to a den-
sity distribution polarized along the vertical (Fig. 4b), or
equivalently θji preferentially around 0◦ or 180◦ (Fig. 1).
This induces a velocity increase along the swimming di-
rection e(cid:107)i (see Eq. (3)). The same is true for the milling
and turning phases (Fig. 4c–d).
Why do individuals tend to swim in-line in the presence
of the fluid? To address this question, we examined the
preferential location of the nearest Voronoi neighbors in
the swarming and schooling phases, in the presence of the
fluid or not (Supplementary Fig. 3). It appears that the
preferential in-line configuration is only present in the full
hydrodynamic model. This is because the side-by-side
configuration becomes unstable when hydrodynamic in-
teractions are considered [40, 42] (Supplementary Fig. 4),
(a) swarming(b) schooling(c) milling(d) turning100r0100r0100r0100r0p/pmaxFig40510alignmentI∥00.51VPM(a)(b)Fig5alignmentIkdipoleintensityIf0.5to the group, for energetic considerations, or when con-
fronted to a danger. Note that the milling and the turn-
ing phases, when the number of swimmers is large, can
break into several smaller groups and thus affect the value
of the order parameters P and M (Supplementary Movies
5-7, and Supplementary Fig. 6).
In summary, we proposed a new model of collective
fish motion that includes behavioral rules and far-field
hydrodynamic interactions. By simulating numerically
the dynamics of this model for a large group of swim-
mers, we showed that, on average, fish swim faster in a
school, due to the presence of the fluid. This suggests
that fish would need less energy to swim in a school for
a given swimming speed. This emergent property results
from the preferential in-line pairing of swimmers, more
robust than the side-by-side configuration. In addition,
we observed a new phase, called the turning phase, which
only exists with the full hydrodynamic model. Finally,
we showed that the fluid has similar effect to the sponta-
neous cognitive rotational noise. These promising results
underline the importance of hydrodynamic interactions
in fish schooling.
In future work, it will be important
to assess the validity of the far-field approximation used
here by integrating the fish wakes into the fluid model.
We are grateful to G. Theraulaz for his valuable
insight. A. F. acknowledges support from A*MIDEX
(ANR-11-IDEX-0001-02), and the Labex MEC (ANR-
10-LABX-0092). E. K. acknowledges sabbatical support
from the Flatiron Institute at the Simons Foundation and
research support from Office of Naval Research (ONR)
through grants N00014-14-1-0421 and N00014-17-1-2287
and the Army Research Office (ARO) through the grant
W911NF-16-1-0074.
∗ [email protected]
[1] C. W. Reynolds, in ACM SIGGRAPH Computer Graph-
ics, Vol. 21 (ACM, 1987) pp. 25–34.
[2] I. D. Couzin and J. Krause, Adv. Study Behav. 32, 1
(2003).
[3] T. Vicsek and A. Zafeiris, Physics Reports 517, 71
(2012).
[4] M. Moussaıd, D. Helbing, and G. Theraulaz, Proc. Nat.
Acad. Sc. USA 108, 6884 (2011).
[5] M. Moussaid, E. G. Guillot, M. Moreau, J. Fehrenbach,
O. Chabiron, S. Lemercier, J. Pettr´e, C. Appert-Rolland,
P. Degond,
and G. Theraulaz, PLoS Comp. Biol. 8,
e1002442 (2012).
[6] E. Shaw, American Scientist 66, 166 (1978).
[7] U. Lopez, J. Gautrais, I. D. Couzin, and G. Theraulaz,
Interface focus (2012), 10.1098/rsfs.2012.0033.
5
[11] I. Aoki, Bull. Jap. Soc. Sci. Fish. 48, 1081 (1982).
[12] A. Huth and C. Wissel, J. Theor. Biol. 156, 365 (1992).
[13] I. D. Couzin, J. Krause, R. James, G. D. Ruxton, and
N. R. Franks, J. Theor. Biol. 218, 1 (2002).
[14] G. ´Odor, Rev. Mod. Phys. 76, 663 (2004).
[15] J. Toner, Y. Tu, and S. Ramaswamy, Ann. Phys. 318,
170 (2005).
[16] M. C. Marchetti, J. Joanny, S. Ramaswamy, T. Liver-
pool, J. Prost, M. Rao, and R. A. Simha, Rev. Mod.
Phys. 85, 1143 (2013).
[17] D. Saintillan and M. J. Shelley, in Complex Fluids in
biological systems (Springer, 2015) pp. 319–355.
[18] A. Huth and C. Wissel, Ecological modelling 75, 135
(1994).
[19] J. Gautrais, F. Ginelli, R. Fournier, S. Blanco, M. So-
ria, H. Chat´e, and G. Theraulaz, PLoS Comput Biol 8,
e1002678 (2012).
[20] D. S. Calovi, U. Lopez, S. Ngo, C. Sire, H. Chat´e, and
G. Theraulaz, New Journal of Physics 16, 015026 (2014).
[21] D. S. Calovi, A. Litchinko, V. Lecheval, U. Lopez,
A. P´erez Escudero, H. Chat´e, C. Sire, and G. Theraulaz,
PLoS Comp. Biol. 14, e1005933 (2018).
[22] J. Krause and G. D. Ruxton, Living in groups (Oxford
University Press, 2002).
[23] D. Weihs, Nature (London) 241, 290 (1973).
[24] S. Alben, in Natural Locomotion in Fluids and on Sur-
faces, The IMA Volumes in Mathematics and its Appli-
cations No. 155, edited by S. Childress, A. Hosoi, W. W.
Schultz, and J. Wang (Springer New York, 2012) pp.
3–13.
[25] A. C. H. Tsang and E. Kanso, J. Nonlin. Sc. 23, 971
(2013).
[26] C. Hemelrijk, D. Reid, H. Hildenbrandt, and J. Padding,
Fish and Fisheries 16, 511 (2015).
[27] J. C. Liao, D. N. Beal, G. V. Lauder, and M. S. Tri-
antafyllou, J. Exp. Biol. 206, 1059 (2003).
[28] J. C. Liao, D. N. Beal, G. V. Lauder, and M. S. Tri-
antafyllou, Science 302, 1566 (2003).
[29] D. N. Beal, F. S. Hover, M. S. Triantafyllou, J. C. Liao,
and G. V. Lauder, J. Fluid Mech. 549, 385 (2006).
[30] H. Bleckmann, Forts. Zool. (1994).
[31] B. L. Partridge and T. J. Pitcher, J. Comp. Physiol. 135,
315 (1980).
[32] K. Faucher, E. Parmentier, C. Becco, N. Vandewalle, and
P. Vandewalle, Animal Behaviour 79, 679 (2010).
[33] S. D. Ryan, B. M. Haines, L. Berlyand, F. Ziebert, and
I. S. Aranson, Physical Review E 83, 050904 (2011).
[34] Inertia is assumed to be negligible, as it was observed in
the experiments made on Kuhlia mugil [19] (α (cid:28) 1 in
Ref. [20]).
[35] J. Lighthill, An informal introduction to theoretical fluid
mechanics (Oxford University Press, New York, NY,
1986).
[36] J. Lighthill, "Mathematical approaches in hydrodynam-
ics," (SIAM, 1991) Chap. Hydrodynamic far fields, pp.
3–20.
[37] A. A. Tchieu, E. Kanso, and P. K. Newton, Proc. R.
Soc. London Ser. A 468, 3006 (2012).
[8] D. V. Radakov, Schooling in the ecology of fish (J. Wiley,
[38] M. Gazzola, A. Tchieu, D. Alexeev, A. de Brauer, and
1973).
P. Koumoutsakos, J. Fluid Mech. 789, 726 (2016).
[9] T. Vicsek, A. Czir´ok, E. Ben-Jacob, I. Cohen,
and
[39] G. V. Lauder and E. G. Drucker, Physiology 17, 235
O. Shochet, Phys. Rev. Lett. 75, 1226 (1995).
(2002).
[10] G. Gr´egoire, H. Chat´e, and Y. Tu, Physica D: Nonlinear
[40] A. D. Becker, H. Masoud, J. W. Newbolt, M. Shelley,
Phenomena 181, 157 (2003).
and L. Ristroph, Nature Communications 6, 8514 (2015).
[41] Following the usage in collective behavior studies, we re-
fer to the different spatial organizations of the swimming
group as "phases", which should not to be confused with
the phases used in equilibrium statistical physics. Here,
they correspond to different regions in the space of or-
der parameters (P , M ), defined with some arbitrariness
when the system is finite.
[42] E. Kanso and A. C. H. Tsang, Fluid Dyn. Res. 46, 061407
(2014).
6
|
1708.07678 | 2 | 1708 | 2018-03-08T10:16:54 | Deciphering mRNA Sequence Determinants of Protein Production Rate | [
"physics.bio-ph",
"cond-mat.stat-mech",
"q-bio.SC"
] | One of the greatest challenges in biophysical models of translation is to identify coding sequences features that affect the rate of translation and therefore the overall protein production in the cell. We propose an analytic method to solve a translation model based on the inhomogeneous totally asymmetric simple exclusion process, which allows us to unveil simple design principles of nucleotide sequences determining protein production rates. Our solution shows an excellent agreement when compared to numerical genome-wide simulations of S. cerevisiae transcript sequences and predicts that the first 10 codons, together with the value of the initiation rate, are the main determinants of protein production rate. Finally, we interpret the obtained analytic results based on the evolutionary role of codons' choice for regulating translation rates and ribosome densities. | physics.bio-ph | physics |
Deciphering mRNA Sequence Determinants of Protein Production Rate
Juraj Szavits-Nossan∗
SUPA, School of Physics and Astronomy, University of Edinburgh,
Peter Guthrie Tait Road, Edinburgh EH9 3FD, United Kingdom
DIMNP UMR 5235, Universit´e de Montpellier and CNRS, F-34095, Montpellier, France and
Laboratoire Charles Coulomb UMR5221, Universit´e de Montpellier and CNRS, F-34095, Montpellier, France
Luca Ciandrini
M. Carmen Romano
SUPA, Institute for Complex Systems and Mathematical Biology,
Department of Physics, Aberdeen AB24 3UE, United Kingdom and
Institute of Medical Sciences, University of Aberdeen,
Foresterhill, Aberdeen AB24 3FX, United Kingdom
One of the greatest challenges in biophysical models of translation is to identify coding sequences
features that affect the rate of translation and therefore the overall protein production in the cell.
We propose an analytic method to solve a translation model based on the inhomogeneous totally
asymmetric simple exclusion process, which allows us to unveil simple design principles of nucleotide
sequences determining protein production rates. Our solution shows an excellent agreement when
compared to numerical genome-wide simulations of S. cerevisiae transcript sequences and predicts
that the first 10 codons, which is the ribosome footprint length on the mRNA, together with the
value of the initiation rate, are the main determinants of protein production rate under physiological
conditions. Finally, we interpret the obtained analytic results based on the evolutionary role of
codons' choice for regulating translation rates and ribosome densities.
PACS numbers: 87.16.aj, 87.10.Mn, 05.60.-k
Translation is one of the major steps in protein biosyn-
thesis. During this process, the nucleotide sequence of a
messenger RNA (mRNA) is translated into a functional
protein. Each nucleotide triplet, called codon, codes for a
specific amino acid, the proteins' building block. There
is experimental evidence that the rate at which a cer-
tain mRNA is translated depends on its specific codon
sequence [1–4], especially in the case of eukaryotes. Iden-
tifying sequence features that determine protein produc-
tion rate, also commonly referred to as translation rate
or efficiency, is a fundamental open question in molecular
biology [1, 3].
Translation is performed by molecular motors called ri-
bosomes, which move unidirectionally along the mRNA.
The amino acids are delivered to the ribosome by
molecules called transfer RNAs (tRNAs), which are spe-
cific to the codon and the amino acid they deliver (Fig. 1).
The dwelling time of a ribosome on a specific codon de-
pends primarily on the abundance of the corresponding
tRNA [5, 6]. Understanding how codon sequences de-
termine protein production rates can potentially unlock
many synthetic applications [7, 8].
The standard biophysical model of translation is
known as the totally asymmetric simple exclusion pro-
cess (TASEP), which captures the concurrent motion of
ribosomes on the mRNA [9, 10].
In this model ribo-
somes progress along the mRNA codon by codon, pro-
vided that the codon a ribosome moves onto is not occu-
pied by another ribosome. The protein production rate
FIG. 1: Sketch of the mRNA translation process involving
initiation (a), elongation consisting of tRNA delivery (b) and
translocation (c), followed by termination (d). We emphasise
that this is an oversimplified scheme of the process and that
actual ribosomes cover (cid:96) = 10 codons. At each elongation
step, the ribosome receives an amino acid from a tRNA that
matches the codon occupied by the A-site of the ribosome (the
"reading" site). After the amino acid is added to the growing
polypeptide, the ribosome translocates one codon forward,
and the process is repeated.
can then be identified as the ribosomal current of this
AUG GUA CUG ACG ... start codonmRNAarate α ld ribosomeA-siteAUG GUA CUG ACG AUU UAU CCU GAC ... UGCtRNAamino acidi-th codonrate kigrowing polypeptidebmRNAA-siteAUG GUA CUG ACG AUU UAU CCU GAC ... mRNAi-th codonrate γ cA-sitegrowing polypeptide... TCT ATA TAT GAA UCU ACA UAAmRNAstop codonrate β polypeptide releaseddA-sitedriven lattice gas. Due to the net current, TASEP is
not in equilibrium and its steady state is only known
for a few special cases [11, 12]. Unfortunately, most
biologically relevant variants of TASEP can be studied
only numerically [13–15], and efficient methods of ex-
ploring a large number of parameters are lacking. The
steady state of the TASEP with non-uniform hopping
rates is a long outstanding problem in nonequilibrium
statistical physics [15, 16]. An analytic prediction of pro-
tein production rates is also needed in order to interpret
recently developed ribosome profiling experiments that
are capable of monitoring ribosome positions along the
mRNA [17]. Such analytical prediction is missing in pre-
vious models of ribosome dynamics implicitly or explic-
itly based on the TASEP [1, 2, 19, 20, 22], and it is key
to decipher sequence determinants of protein production
rates.
In this Letter, we develop a versatile analytic method
to solve TASEP-based models of translation. Our an-
alytic approach, integrated with simulations and experi-
mental data, allows us to efficiently identify the main fea-
tures of mRNA codon sequence that determine the rate of
protein production. The genome-wide comparison of our
analytic predictions with numerical simulations shows an
excellent agreement for the model organism S. cerevisiae
(baker's yeast).
Stochastic model of mRNA translation. We focus on
the model for translation introduced in [1, 23, 24]. The
mRNA is represented by a one-dimensional lattice con-
sisting of L discrete sites (codons), where site 1 designates
the start codon. Ribosomes are represented by particles
that occupy (cid:96) = 10 lattice sites, which is the ribosome
footprint length measured in ribosome profiling experi-
[17]. We identify the position 1 ≤ i ≤ L of a
ments
ribosome with the position of its A-site, which is located
d = 5 lattice sites from the trailing end of the ribosome
(Fig. 1a). A ribosome that waits for a tRNA at position
i is labelled by 1i and a ribosome that has already re-
ceived the correct tRNA and is ready to move is labelled
by 2i. The set of labels of all translating ribosomes on
the lattice is called a configuration C of the system. For
example, C = 11213 denotes a lattice with 2 ribosomes,
one at site 1 waiting for a tRNA and another one at site
13 that has already received the correct tRNA.
Ribosomes initiate translation at rate α by binding to
the mRNA so that their A-site is at the start codon,
provided the sites 1, . . . , (cid:96)− d + 1 are empty (Fig. 1a). A
ribosome at site i makes the transition 1i → 2i at rate ki
dependent on tRNA abundances (Fig. 1b) and moves one
site forward at rate γ (Fig. 1c). Due to steric interactions
between particles, the particle at site i can move only if
there is no ribosome at site i + (cid:96). Termination occurs
when the particle at site L receives the last amino acid,
releases the final protein and detaches from the mRNA,
which we integrate into a single step occurring at rate β
(Fig. 1d).
2
Our main goal is to compute the rate of protein pro-
duction as a function of the parameters of the model,
which are α, β, γ and ki for each of the L codons used.
We assume that translation takes place under steady-
state conditions, so that the ribosomal current is con-
stant along the mRNA and is equal to the rate of protein
production, which in turn is equal to rate at which ri-
bosomes load onto the mRNA and initiate translation.
To this end we define the codon occupation number τi to
be equal to 1 if the i-th codon is occupied by an A-site,
and 0 if otherwise. By definition, the exact steady-state
ribosomal current J then reads
(cid:35)
(1 − τi(C))
P (C),
(1)
(cid:34) (cid:96)(cid:89)
(cid:88)
C
i=1
J = α
where τi(C) denotes the i-th codon occupation number
for configuration C, P (C) denotes the steady-state prob-
ability that the lattice is in configuration C, and the sum-
mation goes over all configurations C. Other quantities of
interest that we compute are the local and total particle
densities ρi = (cid:104)τi(cid:105) and ρ = (1/L)(cid:80)L
i=1 ρi, respectively.
Series expansion method for computing P (C). In order
to find P (C) one has to solve the steady-state master
equation M P = 0, where P is a column vector whose N
elements are the steady-state probabilities P (C) of being
in configuration C, and N denotes the total number of
configurations. The transition rate matrix MC,C(cid:48) is given
by WC(cid:48)→C for C (cid:54)= C(cid:48) and −e(C) for C = C(cid:48), where
WC(cid:48)→C is the transition rate from C(cid:48) to C and e(C) =
C(cid:48)(cid:48)(cid:54)=C WC→C(cid:48)(cid:48) is the total exit rate from C. The exact
solution of the master equation can be formally written
as
(cid:80)
P (C) =
,
(2)
(cid:80)N
detM (p,p)
q=1 detM (q,q)
where p is the position of configuration C in the column
vector P and detM (p,p) is a determinant of the matrix
obtained by removing p-th row and p-th column from M
(see Supplemental Material for details). Unfortunately,
calculating this determinant is feasible only for unrealis-
tically small system sizes.
To circumvent this problem, we exploit the fact that
detM (p,p) is a multivariate polynomial in the variables α,
k1, . . . , kL, β and γ [25]. In the biological literature it is
often assumed that the initiation rate α is a major lim-
iting step of the translation process, mainly determined
by the presence of secondary structures [26, 27]. For this
reason we use α as an expansion parameter and we as-
sume that α (cid:28) γ, β, k1, . . . , kL. By collecting terms with
the same power of α, we can rewrite P (C) as an uni-
variate polynomial f (C) in the variable α with unknown
coefficients fn(C)
f (C)(cid:80)
C f (C)
K(C)(cid:88)
n=0
P (C) =
,
f (C) =
fn(C)αn,
(3)
+
(cid:88)
e0(C) =(cid:80)
C(cid:48)
where the coefficients fn(C) depend on the transition
rates k1, . . . , kL, γ and β (in order to ease the notation,
we leave out the explicit dependence on those parame-
ters). Since we expect that f (C) can be well approxi-
mated by the first few terms, the value of K(C) in Eq. (3
is irrelevant in our study.
In order to find the unknown coefficients fn(C), we
insert Eq. (3) into the master equation M P = 0, col-
lect all the terms with the same n-th power of α and
equate their sum to zero. For a given power n, the re-
sulting equation is similar to the original master equation
in which P (C) is replaced by fn(C) unless the coefficient
multiplying P (C) is α, in which case P (C) is replaced by
fn−1(C). Starting with n = 0, we note the equations for
the coefficients f0(C) have the same form as the original
master equation, but with α = 0, leading to the trivial
solution f0(C) = 0 if C (cid:54)= ∅. Since the terms f (C) are
not normalized, we have the freedom to choose any value
for f0(∅), which we set to 1.
For n ≥ 1, the equation for fn(C) reads
e0(C)fn(C) + fn−1(C)
IC,C(cid:48) =
(1 − IC(cid:48),C)WC(cid:48)→Cfn(C(cid:48)) ,
IC(cid:48),Cfn−1(C(cid:48))
(4)
(cid:88)
C(cid:48)
(cid:88)
C(cid:48)
where IC,C(cid:48) = 1 if WC→C(cid:48) = α and is 0 otherwise, and
C(cid:48)(1−IC,C(cid:48))WC→C(cid:48) is the total exit rate from
C excluding the rate α.
A key observation in our analysis is that fn(C) = 0
whenever the number of particles in C is larger than
n. This follows from the result f0(C) = 1 (0) if C = ∅
(C (cid:54)= ∅) in conjunction with the hierarchical structure
of Eq. (4), which connects configurations differing in the
number of particles by no more than one (a proof for
n = 1 is presented in the Supplemental Material). This
allows us to write Eq. (4) taking into account configura-
tions with only one particle and discarding all the others,
which yields
f1(11) =
f1(1i) =
f1(2i) =
1
k1
γ
ki
ki
γ
f0(∅),
f1(2i−1),
f1(1i),
f1(1L) =
β
kL
2 ≤ i ≤ L,
f1(2L)
1 ≤ i ≤ L − 1.
(5a)
(5b)
(5c)
The solution to Eqs. (5) is given by
f1(1i) =
1
ki
,
f1(2i) =
γ
1
β
i = 1, . . . , L − 1
i = L
(6)
(cid:40) 1
The equations for f2(C) involving configurations with
one and two particles are presented in the Supplemen-
tal Material.
P IPA(C) =
1
ZL
w1(θX(j)),
(13)
Once we determine the coefficients fn(C) up to a de-
sired order n, we can compute the steady-state average
where N (C) is the number of particles in a configuration
C, θ is one of the two particle states 1 and 2, X(i) is
3
of any observable O(C) by inserting P (C) from Eq. (3)
and expanding (cid:104)O(cid:105) around α = 0,
(cid:80)K
(cid:80)
(cid:80)K
(cid:80)
C O(C)fn(C)αn
n=0
C fn(C)αn
n=0
∞(cid:88)
n=0
=
cnαn.
(7)
(cid:104)O(C)(cid:105) =
For example, the first three coefficients c0, c1 and c2 are
given by
c0 =
, c1 =
, c2 =
a1 − c0b1
b0
a2 − c1b1 − c0b2
,
(8)
b0
a0
b0
Notice that the expansion in Eq. (7) is slightly different
for the current J due to an extra α in Eq. (1) and is
n=0 cnαn+1. Using the expressions for
fn(C) and fn(C) computed earlier yields
(cid:88)
C
bn =
where an and bn are defined as
O(C)fn(C),
an =
C
(cid:88)
given by J = (cid:80)∞
(cid:34)
1 − (cid:96)(cid:88)
(cid:18) 1
L(cid:88)
(cid:19)
(cid:18) 1
J = α
1
L
ρ =
i=1
i=1
ki
ρi =
+
ki
1
γi
(cid:19)
1
γi
+
(cid:19)
(cid:18) 1
ki
1
γi
α + O(α2),
+
α + O(α2),
α + O(α2)
,
(10)
(11)
(12)
fn(C).
(9)
(cid:35)
where γi = γ+(β−γ)δi,L. These equations constitute our
main result. Equation (10) shows that if the initiation
rate α is small compared to k1, . . . , kL and γ, than the
protein production rate J depends predominately on the
initiation rate, along with the translocation and elonga-
tion rates of the first 10 codons, corresponding to the ri-
bosome footprint (cid:96). The importance of the first 10 codons
is a direct consequence of the excluded volume interac-
tions: any ribosome already present in that region will
prevent a new ribosome from binding the mRNA. It is
also important to emphasize that (10)–(12) are exact se-
ries expansions around α = 0; the approximation is made
only when the series is truncated.
Independent Particle Approximation (IPA). Interest-
ingly, the excluded volume interactions between particles
have no effect on the first two terms in the series expan-
sion of J. This motivates us to ask how the expansion
in Eq. (10) would look like if we assumed that all parti-
cles are independent, i.e. not experiencing any exclusion
interaction. In our model, the IPA amounts to replacing
P (C) with
N (C)(cid:89)
j=1
4
the position of the i-th particle on the lattice and ZL
is the normalization constant. The weights w1(1i) and
w1(2i) for i = 1, . . . , L are obtained by solving the master
equation for a single particle and are given by α/ki and
α/γi, respectively. The corresponding expressions for the
current J and local density ρi read
α(cid:81)(cid:96)
i=1(1 + pi)
J IPA =
,
ρIPA
i =
pi
1 + pi
,
(14)
where pi = α(1/ki + 1/γi). We will use these results
later in order to determine the importance of ribosome
collisions in real genetic sequences. We also note that
the IPA in (13) provides a good approximation to fn(C)
for n = N (C) when the particles in C are far apart from
each other.
Application to mRNA translation in yeast. We now
apply our results to the transcriptome of S. cerevisiae
using realistic model parameters. The values of α in the
range 0.005 − 4.2 s−1 with the median value of 0.09 s−1
have been previously estimated in Ref. [1] using genome-
wide experimental values of the ribosomal density. We
assume that the rates ki are mainly proportional to the
gene copy number of tRNAs delivering the corresponding
amino acid [28]; the rates are normalized so that the av-
erage codon translation rate is equal to the experimental
value of 10 codons/s [3]; the estimates of all elongation
rates along with the distribution of α is presented in the
Supplemental Material. The translocation rate γ is fixed
to γ = 35 codons/s [29], and termination is assumed to
be fast and comparable to translocation [3], β ≈ γ, so
that γi = γ ∀i in Eqs. (10)-(12).
In total, we analyzed 5836 gene sequences; for each
gene we calculated J, ρ and ρi for 1 ≤ i ≤ L up to and
including the second order of the perturbative expansion
at the corresponding physiological value of α. The re-
sults were then compared to the exact values obtained
numerically with stochastic simulations using the Gille-
spie algorithm [31] by calculating the percent error .
For the protein production rate J, with the zeroth or-
der of the perturbative expansion (predicting that J = α)
we obtain an error of < 5% for only 11% of the genes.
Remarkably, that percentage jumps to 80.7% when the
first-order coefficients in Eq. (6) are taken into account
(Fig. 2).
Including the second-order coefficients (com-
puted numerically from Eq. (4)) does not significantly
improve results, due to a large value of α (cid:39) 0.15 s−1
in about 20% of the genes. Since the coefficients in Eq.
(7) typically alternate in sign, truncating the series will
ultimately lead to a wrong result when the value of α is
large enough. On the other hand, the IPA does not suffer
from this problem and leads to an error < 5% in 94%
of genes, whereby only 1% of genes have > 20%. The
success of the IPA also suggests that ribosome collisions
and traffic jams have a minor effect on the rate of trans-
lation, which is in accordance with recent experimental
evidence [26, 32]. This is also apparent from the density
FIG. 2: (a): Histogram of the percent error measuring the
discrepancy between the protein production rate obtained by
stochastic simulations and the perturbative expansion includ-
ing the independent-particle approximation (IPA). (b) and
(c): Ribosomal density profile obtained by stochastic simula-
tions (black) compared to Eq. (12) (red) for two values of the
initiation rate α, one close to the median value ≈ 0.09 s−1
(gene YAL045C, b) and the other that is close to the 90th
percentile value ≈ 0.23 s−1 (gene YGL034C, c).
profile ρi, which is very well approximated by the linear
approximation in Eq. (12), even for larger values of α
(Fig. 2 b and c).
Identifying determinants of mRNA translation. Our
analytic prediction allows us to decompose the contribu-
tions from initiation and elongation to the rate of trans-
lation J, thereby addressing a long-standing question
about main determinants of protein production rate. Re-
markably, the expressions for the current obtained with
both the first-order and the independent particle approx-
imation involve only the first (cid:96) = 10 codons. This result
therefore strongly indicates that, together with the ini-
tiation rate, the first 10 codons of the mRNA are the
key determinants of the protein production rate by pre-
venting a new ribosome from binding which effectively
decreases the initiation rate.
If we assume that the cell maximizes the rate of pro-
tein production, given the above result, we would ex-
n=0,Taylorn=1,Taylorn=2,TaylorIPA0-55-1010-1515-20>20020406080100ϵ[%]percentageofgenes[%]a0204060801000.050.100.150.20i ib0204060801001200.050.100.150.20i ic5
Acknowledgments. MCR and LC contributed equally
to this work. JSN was supported by the Leverhulme
Trust Early Career Fellowship. MCR was supported
by the Biotechnology and Biological Sciences Research
Council (BBSRC) BB/N017161/1 and the Scottish Uni-
versities Life Sciences Alliance. LC thanks the CNRS for
having granted him a demi-d´el´egation (2017–18).
∗ Electronic address: [email protected]
[1] H. Gingold, and Y. Pilpel, Mol. Syst. Biol. 7, 481 (2011).
[2] A.J. Kemp, R. Betney, L. Ciandrini, A. C. Schwenger, M.
C. Romano, and I. Stansfield, Mol. Microbiol. 87(2), 284
(2013).
[3] D. Chu, E. Kazana, N. Bellanger, T. Singh, M. F. Tuite,
and T. von der Haar, EMBO J. 33(1), 21 (2014).
[4] B. Gorgoni, L. Ciandrini, M. R. McFarland, M. C. Ro-
mano, and I. Stansfield, Nucleic Acids Res. 44(19), 9231
(2016).
[5] S. Varenne, J. Buc, R. Lloubes, and C. Lazdunski, J. Mol
Biol. 180(3), 549 (1984).
[6] M. A. Sorensen, C. G. Kurland, and S. Pedersen, J. Mol.
Biol. 207, 365–377 (1989).
[7] T. E. Gorochowski, I. Avcilar-Kucukgoze, R. A. Boven-
berg, J.A. Roubos, and Z. Ignatova, ACS Synth. Biol.
5(7), 710 (2016).
[8] G. Boel et al, Nature 529(7586), 358 (2016).
[9] C. T. MacDonald, J. H. Gibbs, and A. C. Pipkin, Biopoly-
mers 6, 1–25 (1968); C. T. MacDonald and J. H. Gibbs,
Biopolymers 7, 707–25 (1969).
[10] H. Zur, and T. Tuller, Nucleic Acids Res. 44(19), 9031–
9049 (2016).
[11] B. Derrida, M. R. Evans, V. Hakim, and V. Pasquier, J.
Phys. A: Math. Gen. 26, 1493 (1993); G. M. Schutz, and
E. Domany, J. Stat. Phys. 72, 277–96 (1993).
[12] R. A. Blythe, and M. R. Evans, J. Phys. A: Math. Theor.
40, R333–R441 (2007).
[13] T. Chou, and G. Lakatos, Phys. Rev. Lett. 93, 198101
(2004).
[14] L. B. Shaw, J. P. Sethna, and K. H. Lee, Phys. Rev. E
70, 021901 (2004).
[15] R. K. P. Zia, J. J. Dong, and B. Schmittmann, J. Stat.
Phys. 144, 405 (2011).
[16] J. Schmidt, V. Popkov and A. Schadschneider, EPL 110,
20008 (2015)
[17] N. T. Ingolia, S. Ghaemmaghami, J. R. S. Newman, and
J. S. Weissman, Science 324, 218–223 (2009).
[18] M. A. Gilchrist, A. Wagner, J. Theor. Biol., 239(4), 417-
434 (2006).
[19] N. Mitarai, K. Sneppen, and S. Pedersen, J. Mol. Biol.,
382(1), 236-245 (2008).
[20] S. Reuveni, I. Meilijson, M. Kupiec, E. Ruppin, and T.
Tuller, PLoS Comput. Biol. 7, e1002127 (2011).
[21] L. Ciandrini, I. Stansfield, and M. C. Romano, PLoS
Comput. Biol. 9(1), e1002866 (2013).
[22] P. Shah, Y. Ding, M. Niemczyk, G. Kudla, and J.B.
Plotkin, Cell 153(7), 1589 (2013).
[23] L. Ciandrini, I. Stansfield, and M. C. Romano. Phys. Rev.
E 81, 051904 (2010).
[24] S. Klumpp, Y. Chai, and R. Lipowsky, Phys. Rev. E 78,
FIG. 3: Histogram of ηJ (yellow) and ηρ (blue) for the S.
cerevisiae genome.
pect to find a signature in the genome for selecting ef-
ficient fast codons at the beginning of each gene. We
test this hypothesis by computing the ribosomal current
for the fastest (J F ) and the slowest (J S) set of first
(cid:96) synonymous codons for each gene, using the IPA. A
score ηJ = (J − J S)/(J F − J S) is then assigned to each
gene, which is 1 (0) when the sequence corresponds to
the fastest (slowest) codon sequence. On the other hand,
one might assume that the cell not only tries to maximize
protein production rates, but at the same time it tries
to minimize the ribosome density ρ on mRNAs. This as-
sumption is motivated by the fact that ribosomes are lim-
iting [22, 33] and highly costly in terms of cellular energy
resources [34], and therefore ribosome queues are to be
avoided. Hence, we also compute ηρ = (ρ−ρS)/(ρF −ρS)
for each gene, where ρF and ρS denote the ribosome den-
sity for the fastest and slowest set of synonymous codons,
respectively.
Figure 3 shows the histogram of ηJ (yellow) and ηρ
(blue) computed for 5836 genes of S.cerevisiae. Both his-
tograms show an average of 0.7 suggesting the selection
of fast codons near the start codon to maximize J, as well
as an overall selection of fast codons along the mRNA to
minimize ρ. However, the width of the distribution of ηρ
is substantially smaller than the one of ηJ . This might
indicate that the optimization of protein production rate
is strongly dependent on the particular gene, since dif-
ferent proteins are needed at different concentrations. In
contrast, the minimization of the number of ribosomes
on mRNAs could be a more general constraint.
Conclusions. We have presented an analytic method
that allows us to quantify sequence determinants of pro-
tein production rates, using a model for translation that
is based on an inhomogeneous exclusion process. Our
results demonstrate that the rate of protein production
is largely determined by the initiation rate and the elon-
gation rates of the first 10 codons (the ribosome foot-
print length on the mRNA), which control how fast ribo-
somes load onto the mRNA, whereby ribosome collisions
and queues have a minor effect under physiological con-
ditions.
041909 (2008).
(2003).
[25] J. Szavits-Nossan, J. Phys. A: Math. Theor. 46, 315001
[30] Y. Arava, Y. Wang, J. D. Storey, C. L. Liu, P. O. Brown,
(2013).
[26] G. Kudla, A. W. Murray, D. Tollervey, and J. B. Plotkin,
Science 324, 255–258 (2009).
[27] H. M. Salis, E. A. Mirsky, and C. A. Voigt, Nature
Biotechnol. 27, 946–950 (2009).
and D. Herschlag, PNAS 100(7), 3889–3894 (2003).
[31] D. T. Gillespie, J. Phys. Chem. 81(25), 2340 (1977).
[32] B. S. Laursen, H. P. Sørensen, K. K. Mortensen, and H.
U. Sperling-Petersen, Microbiol. Mol. Biol. Rev. 69, 101
(2005).
[28] R. Percudani, A. Pavesi, and S. Ottonello, J. Mol. Biol.
[33] P. Greulich, L. Ciandrini, R. J. Allen, and M. C. Romano.
268(2), 322–330 (1997).
[29] A. Savelsbergh, V. I. Katunin, D. Mohr, F. Peske, M. V.
Rodnina, and W. Wintermeyer, Mol. Cell 11, 1517–1523
Phys. Rev. E 85, 011142 (2012).
[34] K. C. Keiler, Nature Rev. Microbiol. 13, 285 (2015).
6
Supplemental Material to:
Deciphering mRNA Sequence Determinants of
Protein Production Rate
7
EXACT SOLUTION OF THE STEADY-STATE MASTER EQUATION
The exact solution, Eq. (3) in the main text, pertains to any ergodic Markov jump process with a finite number of
states. In order to derive the exact solution, let us write again the steady-state master equation in a matrix form
(S1)
where P is a column vector whose elements are steady-state probabilities Pi, the N states are indexed by i = 1 . . . ,N
and the N × N matrix M is given by
M P = 0,
Mij =
k(cid:54)=i Wik
i (cid:54)= j
i = j,
(S2)
where Wij is a transition rate from a state i to j. We note from Eq. (S2) that the sum of all elements in each column
of M is zero
N(cid:88)
N(cid:88)
Mij =
Wji −
Wjk = 0.
(S3)
Consequently, the sum all row vectors of M is zero, which means that the vectors are linearly dependent and thus the
determinant of M is equal to zero,
i=1
i=1
i(cid:54)=j
k=1
k(cid:54)=j
Wjk
(cid:40)
−(cid:80)
N(cid:88)
Combining Eq. (S4) with the Laplace expansion of a determinant yields
detM = 0.
N(cid:88)
N(cid:88)
0 = detM =
MijCij =
MijCij.
i=1
j=1
Here Cij is a cofactor of M defined as Cij = (−1)i+jdetM (i,j), where the matrix M (i,j) is obtained from M by
removing i-th row and j-th column . Inserting Eq. (S2) into (S5) gives
N(cid:88)
N(cid:88)
i=1
i(cid:54)=j
detM =
=
MijCij + MjjCjj =
Mij(Cij − Cjj) = 0,
N(cid:88)
N(cid:88)
MijCij −
MkjCjj
i=1
i(cid:54)=j
k=1
k(cid:54)=j
(S4)
(S5)
(S6)
(S7)
(S8)
(S9)
from which we conclude that
i=1
i(cid:54)=j
Cij = Cjj,
for any i and j, i.e. the cofactor Cij does not depend on the state i. Inserting Eq. (S7) back into Eq. (S4) gives
N(cid:88)
j=1
MijCjj = 0,
which is precisely the starting steady-state master equation, Eq. (S1). We thus conclude that
Cii(cid:80)
j Cjj
(cid:80)N
=
detM (i,i)
j=1 detM (j,j)
.
Pi =
ARGUMENT FOR f (n)(C) = 0 WHENEVER THE NUMBER OF PARTICLES IN
A CONFIGURATION C IS LARGER THAN n
8
As noted in the main text, a key observation in our analysis is that f (n)(C) = 0 whenever the number of particles
in C is larger than n. We present here the argument for the case of n = 1; a similar argument applies to higher orders.
For n = 1, using Eq. (5) in the main text we obtain
(cid:21)
(cid:20)(cid:88)
C(cid:48)
f (1)(C) =
1
e(C)
(cid:88)
C(cid:48)
IC(cid:48)→CδC(cid:48),∅ +
(1 − IC(cid:48)→C)WC(cid:48)→Cf (1)(C(cid:48)) − δC,∅
.
(S10)
Consider C to be one full lattice configuration with all N = (cid:98)(L − 1 + (cid:96))/(cid:96)(cid:99) particles in state 1. The first term on
the right hand side of Eq. (S10) is zero because there is no direct transition from the empty configuration ∅ to C.
The second term on the right hand side is also zero, because C can only be accessed through the transition rate α.
Finally, the third term on the right hand side is clearly zero. Hence, f (1)(C) = 0. Now, it is easily seen that all full
lattice configurations C(cid:48) that can be accessed only from C will fulfil f (1)(C(cid:48)) = 0. Iterating further this procedure
to the configurations that can be accessed from those, it becomes clear that all full lattice configurations C have
f (1)(C) = 0. This argument can be further extended to all configurations with N − 1, N − 2, N − 3, ..., 2 particles,
resulting in f (1)(C) = 0 for all of them.
The situation changes when we consider configurations with only 1 particle on the lattice, since then there is a
direct transition from the empty configuration ∅ to one of those configurations, leading to f (1)(C) (cid:54)= 0. The rest of the
coefficients f (1)(C) for 1-particle configurations will depend on each other, and hence, they do not vanish. Therefore,
we can conclude that only 1-particle configurations have non-vanishing f (1)(C). For n = 1, this key observation allows
us to write Eq. (5) taking into account configurations with only one particle and discarding all the others, which
yields Eqs. (6a)-(6c) in the main text.
EQUATIONS FOR THE SECOND-ORDER COEFFICIENTS f (2)(C)
As stated in the main text, all second-order coefficients f (2)(C) whereby a configuration C has more than two
particles are equal to zero. The equations for the non-zero coefficients f (2)(C) are presented below, where we use the
the notation γj = γ for 1 ≤ j ≤ L − 1 and γL = β.
First, we look at configurations with one particle at site i = 1 and the other at site j = i + l, . . . , L:
where we used the Kronecker delta to account for the excluded volume interaction. Next, we look at configurations
with particles at sites i = 2, . . . , L − l and j = i + l, . . . , L.
(cid:104)
(cid:104)
(cid:104)
(cid:104)
(cid:104)
(cid:104)
(cid:104)
(cid:104)
f (1)(1j) + (1 − δ1+l,j)γj−1f (2)(112j−1)
(cid:105)
f (1)(2j) + kjf (2)(111j)
f (2)(111j) =
f (2)(112j) =
f (2)(211j) =
f (2)(212j) =
1
k1 + kj
1
1
k1 + γj
(1 − δ1+l,j)γ1 + kj
(1 − δ1+l,j)γ1 + γj
1
k1f (2)(111j) + (1 − δ1+l,j)γj−1f (2)(212j−1)
k1f (2)(112j) + kjf (2)(211j)
.
f (2)(1i1j) =
f (2)(1i2j) =
f (2)(2i1j) =
f (2)(2i2j) =
1
ki + kj
1
1
ki + γj
(1 − δi+l,j)γi + kj
(1 − δi+l,j)γi + γj
1
γi−1f (2)(2i−11j) + (1 − δi+l,j)γj−1f (2)(1i2j−1)
(cid:105)
γi−1f (2)(2i−12j) + kjf (2)(1i1j)
kif (2)(1i1j) + (1 − δi+l,j)γj−1f (2)(2i2j−1)
kif (2)(1i2j) + kjf (2)(2i1j)
.
(cid:105)
(cid:105)
(cid:105)
(cid:105)
(cid:105)
(cid:105)
(S11a)
(S11b)
(S11c)
(S11d)
(S12a)
(S12b)
(S12c)
(S12d)
Finally, the equations for configurations with only one particle are given by
(cid:105)
γLf (2)(112L) + f (1)(∅)
,
γL
γ1
f (2)(212L),
f (2)(11) +
γi−1f (2)(2i−1) + θ(L − l + 1 − i)γLf (2)(1i2L) − f (1)(1i)
kif (2)(1i) + θ(L − l + 1 − i)γLf (2)(2i2L) − f (1)(2i)
(cid:105)
,
(cid:104)
(cid:104)
(cid:104)
1
k1
k1
γ1
1
ki
1
γi
f (2)(11) =
f (2)(21) =
f (2)(1i) =
f (2)(2i) =
9
(S13a)
(S13b)
,
i = 2, . . . , L
(S13c)
(cid:105)
i = 2, . . . , L
(S13d)
where θ(n) = 0 for n < 0 and 1 for n ≥ 0.
The equations (S11)-(S13) can be easily solved numerically by iteration. We first solve Eqs. (S11) for i = 1, j = l+1.
We then iterate the recursion relation in Eqs. (S11) for fixed i = 1 and j = 2 + l, . . . , L. We then solve equations
(S12) for fixed i = 2 and j = l + 3, . . . , L and repeat this procedure until i = L− l and j = L. Finally, we calculate the
coefficients f (2)(1i) and f (2)(2i) for the single-particle configurations using f (2)(1i2L) and f (2)(2i2L) that we solved
in the previous steps. We note that the only unknown that we cannot determine is f (1)(∅), but it turns out that all
terms containing f (1)(∅) will cancel out later when we calculate the coefficients in the series expansions (11)-(13) in
the main text.
ESTIMATES OF THE CODON ELONGATION RATES
The total translation rate ωi of the codon i is:
(S14)
where the translocation rate γ is set to 35s−1 as mentioned in the main text. The codon elongation rate ki, which
represents the average arrival and recognition time of the cognate tRNA, is determined by following the procedure
introduced in [1], which we report here.
+
=
,
1
ki
1
γ
1
ωi
For each of the 41 tRNAs types j we consider their gene copy number (GCN), which allows us to provide a first
estimate of the rate kj:
(cid:80)41
GCNj
j=1 GCNj
,
kj = r
(S15)
where GCNj is the gene copy number of the tRNA of type j with j = 1, . . . , 41, and r is a proportionality constant.
These rates were then adjusted to take into account experimental evidence suggesting that the translation rates of
codons using the G-U wobble are reduced by 39% compared to their G-C counterparts; analogously, codons using the
wobble I-C and codons using the wobble I-A are reduced by 36% relative to their I-U counterparts [2]. To calculate
the proportionality constant r, we used the experimental value of 10 codons/s for the average codon translation rate
(cid:104)ωi(cid:105) defined as
61(cid:88)
(cid:18) kiγ
(cid:19) ni
ki + γ
n
i=1
(cid:104)ωi(cid:105) =
(S16)
of exactly type i, and n = (cid:80)61
elongation rates ki for 61 codons.
where ni/n is the relative abundance of all different codon types in the cell, ni is total number of codons in the cell
i=1 ni (the three STOP codons are excluded) [1]. Table I summarises the resulting
The estimates provided are valid in physiological conditions, when amino acids are not limiting. In case of amino
acid starvation the rates of the codons affected should be modified (the elongation rates depend on the abundance of
tRNAs charged with the correct amino acid).
DISTRIBUTION OF THE TRANSLATION INITIATION RATES
The translation initiation rates α were estimated for each gene in Ref. [1] by matching the average ribosome density
predicted by the model with the experimental value of the ribosome density obtained in Ref. [3]. The resulting values
10
tRNA anti-codon codon GCN ki [1/s]
Ala1
Ala1
Ala2
Ala2
Arg1
Arg2
Arg2
Arg2
Arg3
Arg4
Asn
Asn
Asp
Asp
Cys
Cys
Gln1
Gln2
Glu3
Glu4
Gly1
Gly1
Gly2
Gly3
His
His
Ile1
Ile2
Ile2
Leu1
Leu1
IGC
IGC
UGC
UGC
CCU
ICG
ICG
ICG
UCU
CCG
GUU
GUU
GUC
GUC
GCA
GCA
UUG
CUG
UUC
CUC
GCC
GCC
UCC
CCC
GUG
GUG
UAU
IAU
IAU
UAG
UAG
GCU 11.00 18.34
GCC 11.00 11.74
8.34
GCA 5.00
8.34
GCG 5.00
AGG 1.00
1.67
10.01
CGU 6.00
6.40
CGC 6.00
CGA 6.00
6.40
AGA 11.00 18.34
CGG 1.00
1.67
AAU 10.00 10.17
AAC 10.00 16.68
GAU 15.00 15.26
GAC 15.00 25.01
4.07
UGU 4.00
6.67
UGC 4.00
CAA 9.00
15.01
CAG 1.00
1.67
GAA 14.00 23.35
GAG 2.00
3.34
GGU 16.00 16.28
GGC 16.00 26.68
5.00
GGA 3.00
3.34
GGG 2.00
CAU 7.00
7.12
11.67
CAC 7.00
AUA 2.00
3.34
AUU 13.00 21.68
AUC 13.00 13.87
5.00
CUA 3.00
CUG 3.00
5.00
tRNA anti-codon codon GCN ki [1/s]
Leu3
Leu4
Leu5
Leu5
Lys1
Lys2
Met
Phe
Phe
Pro1
Pro1
Pro2
Pro2
Ser2
Ser2
Ser3
Ser3
Ser4
Ser5
Thr1
Thr1
Thr2
Thr3
Trp
Tyr
Tyr
Val1
Val1
Val2
Val2b
CAA
UAA
GAG
GAG
CUU
UUU
CAU
GAA
GAA
UGG
UGG
IGG
IGG
IGA
IGA
GCU
GCU
UGA
CGA
IGU
IGU
CGU
UGU
CCA
GIA
GIA
IAC
IAC
UAC
CAC
UUG 10.00 16.68
11.67
UUA 7.00
1.02
CUU 1.00
CUC 1.00
1.67
AAG 14.00 23.35
11.67
AAA 7.00
AUG 5.00
8.34
UUU 10.00 10.17
UUC 10.00 16.68
CCA 10.00 16.68
CCG 10.00 16.68
CCU 2.00
3.34
CCC 2.00
2.13
UCU 11.00 18.34
UCC 11.00 11.74
4.07
AGU 4.00
AGC 4.00
6.67
5.00
UCA 3.00
UCG 1.00
1.67
ACU 11.00 18.34
ACC 11.00 11.74
1.67
ACG 1.00
6.67
ACA 4.00
UGG 6.00
10.01
8.14
UAU 8.00
UAC 8.00
13.34
GUU 14.00 23.35
GUC 14.00 14.94
3.34
GUA 2.00
GUG 2.00
3.34
TABLE I: Elongation rates ki considering supply (gene copy number of tRNAs) and wobble base-pairing.
for 5836 genes that we analyzed in the main text are in the range 0.005 − 4.2 s−1 with the median value of 0.09 s−1.
The histogram of the rates is presented in Figure S1, showing that the initiation rate α for the majority of genes is
indeed much smaller than the smallest elongation rate, which has the value of 1.02 codons/s for the CUU codon (see
Table I). The list of all translation initiation rates can be found in the Supplemental Material of Ref. [1].
∗ Electronic address: [email protected]
[1] L. Ciandrini, I. Stansfield, and M. C. Romano, PLoS Comput. Biol. 9(1), e1002866 (2013).
[2] M. A. Gilchrist and A. Wagner, J. Theor. Biol. 239 417–434 (2006).
[3] Y. Arava, Y. Wang, J. D. Storey, C. L. Liu, P. O. Brown, and D. Herschlag, PNAS 100(7), 3889–3894 (2003).
11
FIG. S1: Histogram of the estimated initiation rates for 5836 genes from the S. cerevisiae genome. The 95th percentile value of
α is ≈ 0.3, which is much smaller that the smallest elongation rate, which is 1.02 codons/s for the CUU codon (vertical dashed
line). Only 15 genes have α larger than mini{ki} = 1.02 codons/s.
0.00.20.40.60.81.01.20200400600800translationinitiationrateα[1/s]numberofgeneskCUU |
1202.6108 | 2 | 1202 | 2012-04-25T03:14:14 | Avian magnetoreception model realized by coupling magnetite-based mechanism with radical-pair-based mechanism | [
"physics.bio-ph"
] | Many animal species were verified to use geomagnetic field for their navigation, but the biophysical mechanism of magnetoreception has remained enigmatic. This paper presents a special biophysical model that consists of magnetite-based and radical-pair-based mechanisms for avian magnetoreception. The amplitude of the resultant magnetic field around the magnetic particles corresponds to the geomagnetic field direction and affects the yield of singlet/triplet state products in the radical-pair reactions. Therefore, in the proposed model, the singlet/triplet state product yields are related to the geomagnetic field information for orientational detection. The resultant magnetic fields corresponding to two materials with different magnetic properties were analyzed under different geomagnetic field directions. The results showed that ferromagnetic particles in organisms can provide more significant changes in singlet state products than superparamagnetic particles, and the period of variation for the singlet state products with an included angle in the geomagnetic field is approximately 180{\deg} when the magnetic particles are ferromagnetic materials, consistent with the experimental results obtained from avian magnetic compass. Further, the calculated results of the singlet state products in a reception plane showed that the proposed model can explain the avian magnetoreception mechanism with an inclination compass. | physics.bio-ph | physics | 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
Avian magnetoreception model realized by
coupling magnetite-based mechanism with
radical-pair-based mechanism
Yan Lu1 and Tao Song1,2*
1Beijing Key Laboratory of Bioelectromagnetism, Institute of Electrical Engineering, Chinese Academy of
Sciences, P. O. Box 2703, Beijing 100190, P. R. China
2France-China Bio-Mineralization and Nano-Structures Laboratory (BioMNSL), P. O. Box 2703, Beijing
100190, P. R. China
*Corresponding author ([email protected])
ABSTRACT
Many animal species were verified to use geomagnetic field for their navigation, but the
biophysical mechanism of magnetoreception has remained enigmatic. This paper presents a
special biophysical model
that consists of magnetite-based and radical-pair-based
mechanisms for avian magnetoreception. The amplitude of the resultant magnetic field
around the magnetic particles corresponds to the geomagnetic field direction and affects the
yield of singlet/triplet state products in the radical-pair reactions. Therefore, in the proposed
model, the singlet/triplet state product yields are related to the geomagnetic field information
for orientational detection. The resultant magnetic fields corresponding to two materials with
different magnetic properties were analyzed under different geomagnetic field directions. The
results showed that ferromagnetic particles in organisms can provide more significant
changes in singlet state products than superparamagnetic particles, and the period of variation
for the singlet state products with an included angle in the geomagnetic field is approximately
180° when the magnetic particles are ferromagnetic materials, consistent with the
experimental results obtained from avian magnetic compass. Further, the calculated results of
the singlet state products in a reception plane showed that the proposed model can explain the
avian magnetoreception mechanism with an inclination compass.
Keywords: magnetoreception, orientation, geomagnetic field, magnetic particles, radical
pair
Ⅰ. INTRODUCTION
Various animals use geomagnetic field for their navigation. The use of magnetic compass by
migratory birds was first described in European robins [1]. In the following years, it has been
further demonstrated in more than 50 species [2]. Recently, two most convincing biophysical
mechanisms of magnetoreception in birds have been proposed: magnetite-based mechanism
[3–5] and radical pair-based mechanism [6–8].
The magnetite-based mechanism is established on the discoveries that magnetic particles are
widely found in bacteria and higher organisms [9, 10]. These magnetic particles were
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
suggested to be involved in magnetoreception, verified by various biophysical methods.
Some magnetite-based models have been presented and theoretically evaluated in efforts to
explain magnetoreception [11–15].
The model proposed by Kirschvink et al. [11, 12] was based on the presumption that the
magnetic particles in organisms are single-domain (SD) magnetites. The authors believed that
external magnetic field exerts a magnetic torque on the particle, which rotates the particle to
open or close the ion channel and produces nerve signals. The model presented by Davila et
al. [13] was based on superparamagnetic (SP) nanoparticles. They assumed that SP
nanoparticles can be magnetized, and the local magnetic field in the cell is amplified by
orders of magnitude under external magnetic field. Thus, magnetic particles experience an
attractive or repulsive force to induce their displacement, which can induce primary receptor
potential via strain-sensitive membrane channels to create nerve signals. Fleissner et al. [14],
and Solov’yov et al. [15] analyzed the magnetoreception process in birds using two types of
ferromagnetic magnetite particles based on the model of Davila et al. [13]. All the models
mentioned above indicated that nerve signals are transmitted through the ophthalmic branch
of the trigeminal nerve to the brain, and magnetic fields are used as compass or component of
a navigational “map” for the navigation of birds [16, 17].
However, these theories cannot explain the light dependence of magnetic orientation
confirmed by several behavioral experiments [18–20]. Further, Zapka et al. [21] reported that
European robins with bilateral section in the ophthalmic branch of the trigeminal nerve can
still use their magnetic compass for orientation, in conflict with the magnetite-based
mechanism because the mechanism requires trigeminal nerve signal transmission.
On the other hand, the radical-pair-based mechanism is established on the proposition that
weak magnetic fields can affect free-radical recombination reactions [22–24]. The
singlet/triplet radical-pair product can act as the receptor of the external magnetic field, and
migratory birds can use it to sense changes in the geomagnetic field. The magnetosensitivity
of the radical-pair reactions has to be anisotropic to detect the direction information of the
geomagnetic field using the radical-pair mechanism. Theoretically, the molecules that provide
the radical pairs can possess the required properties to produce anisotropic effect [25, 26].
Thus, Ritz et al. [27] proposed a special structure of radical pairs called "reference-probe"
design, where the internal magnetic field of the "reference" radical should be much higher
than the external field, but the "probe" radical has very small or no internal magnetic field.
Lau et al. [28] suggested that the molecules that play host to the magnetically sensitive
radical-pair intermediates must be immobilized and rotationally ordered within receptor cells.
Then, a partially rotational disorder can cause anisotropic responses of differently oriented
radical pairs within the same cell. Although anisotropic radical-pair magnetic field effect has
been observed in several solutions or liquid crystal experiments [29–34], no direct evidence
in terms of microcosmic mechanism has been obtained until now.
In addition, some research works observed radical pairs subjected to magnetic fields
produced by iron-containing particles and structures [35–39]. Binhi et al. [35] speculated that
the rate of intracellular free-radical biochemical reactions will be altered by stray fields
produced in different orientations by the intracellular magnet, thought of as an indirect
torque-transduction pathway for magnetoreception by Winklhofer et al. [40], who believed
that the model can explain the conflicting results reported by Zapka [21]. To achieve
2
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
maximum angular sensitivity, the Binhi compass model requires that the intracellular magnet
should rotate, and the magnet should be coupled to a soft elastic matrix. Cohen et al. [37, 38]
demonstrated how magnetic nanostructures can catalyze intersystem crossings in molecular
radical pairs. The radical pair was assumed to be randomly distributed and randomly oriented
on the surface of the magnetic nanostructure. Cohen et al. modeled the nanostructure as a
uniformly magnetized sphere to calculate the magnetic field gradient at the surface of the
sphere. The effect of the geomagnetic field on the magnetic field gradient near the magnetic
nanostructures was not considered. Cai [39] demonstrated how to optimize the design of a
chemical compass using a much better directional sensitivity by simply employing a gradient
field created in the vicinity of a hard ferromagnetic nanostructure. He assigned directly the
value of the gradient field to analyze the magnetic field sensitivity of the chemical compass
without considering the real structure of the nanoparticles in bird.
In this paper, the equivalent model of magnetic particles was established based on the real
nanoparticle structure
in birds. A special combination of magnetite-based and
radical-pair-based mechanisms was analyzed to evaluate the justifiability of the model. In
contrast to existing models, the amplitude of the resultant magnetic field around the magnetic
particles (but not the magnetic field gradient) was adopted to represent the geomagnetic field
direction. Based on the amplitude variation of the resultant magnetic field, the singlet state
product yield of the reactions was calculated using the radical-pair mechanism. Thus, product
yield is related to the geomagnetic field information.
Ⅱ. COMBINED MECHANISM
The subtle iron-containing structure in the skin of the upper beak of birds has three
subcellular components containing iron: chains of maghemite crystals, magnetite clusters,
and iron-coated vesicle [14]. The equivalent model of the magnetic particles in birds was
based on this structure. The magnetic field induced by the magnetic particles and the
geomagnetic field produce a magnetic resultant field, whose amplitude depends on the spatial
location and the direction of the geomagnetic field. Hence, the directional change in the
geomagnetic field can be represented by the resultant magnetic field amplitude.
According to the radical-pair mechanism, the radical pair D++A- is created by an electron
transferred from donor molecule D to acceptor molecule A. The spin directions of radical pair
D++A- have an anti-parallel alignment, and the spin state is called singlet state; when the
alignment is parallel, it is called a "triplet state." The radical pair originates from a singlet
state; its singlet and triplet states are interconverted by hyperfine interaction, and the process
is affected by the applied magnetic field. Therefore, the product yield of a singlet/triplet state
varies with the change in the resultant magnetic field arising from the change in the
geomagnetic field direction. The radical-pair product yield is thus related to the geomagnetic
field information. Consequently, alteration in the product yield will change the number of
neurotransmitters and results in an increase or decrease of the signal in the nerve cell that
receives the neurotransmitters for navigation. To explain the transmission process,
cryptochrome was suggested as a potential primary magnetoreceptor to receive the
information of the change in the radical-pair product yields [7].
the separate disadvantages of
The proposed mechanism can well overcome
the
3
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
magnetite-based and radical-pair-based mechanisms. Dealing with the role of the magnetic
particles in changing the direction information of the geomagnetic field to amplitude of the
resultant magnetic field is the core of the proposed model. Past related research has not
considered this option. Furthermore, because the equivalent model of magnetic particles was
established based on real bird nanoparticle structures, if more convictive experimental
evidence for anisotropic interaction of radical pairs are obtained from future works, this
mechanism can also be used to analyze the effect of magnetic particles, which exist
universally in migrant birds [41], on the radical-pair product yield.
Ⅲ. CALCULATION MODEL
To analyze the resultant magnetic field produced by the interaction between the magnetic
particles and the geomagnetic field, the dendrites in the upper beak that contain magnetic
particles were simulated as a cylinder. The cylinder length was designated as l=300 μm, and
the diameter was designated as D=10 μm, by evaluating the size of the magnetic particles and
dendrites [10], as shown in Fig. 1(a).
Z
e2
d2
e1
R = 0 . 4 m m
o’
O
a2
curve l2
b2
o’’
d=0.4m
m
a1
b1
curve l1
d1
c
X
reception plane S
3
2
1
0
-1
-2
-3
-1600
]
T
[
B
Y
SP Particles
Hc=0
Ms=480kA/m
-50
]
T
[
B
0.25
0
-0.25
50
H [kA/m]
-800
0
H [kA/m]
800
1600
(a) (b)
Fig. 1 Magnetic field calculation.
(a) Three-dimensional model for magnetic particles. The cylinder is used to simulate the magnetic particles in
birds. The centre of the cylinder is the origin of the coordinate system for calculation. (b) Experimental B–H
curve of the SP nanoparticles measured by VSM.
The x-axis is defined as the central axis of the cylinder, the y-axis is perpendicular to the
central axis of the cylinder, and the z-axis is perpendicular to the xy plane and parallel to the
left and right surfaces of the cylinder. This rough assumption is not suitable for quantitative
calculation because of insufficient biophysical experimental data, but it is sufficient for
qualitative analysis of the effects of the resultant magnetic field on the radical pair product
yield.
4
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
To investigate the resultant magnetic field distribution, curve l1 in the xy plane and curve l2 in
the xz plane, shown in Fig. 1(a), were chosen to calculate the amplitude of the resultant
magnetic field at different positions when the applied magnetic field changes. a1–b1 and a2–b2
and d1–e1 and d2–e2 are line segments parallel to the x-axis; the distance from o' (centerpoint
of the left surface of the cylinder) to the line is ±0.4 mm. b1–d1 and b2–d2 are arcs with a
radius of 0.4 mm and an angle of 180°, whose centerpoint is o'.
For convenience of calculation, the bird’s body was assumed to be parallel to the x-axis of the
reference coordinate system. Therefore, when the bird changes its angle of flight attitude, the
direction of the geomagnetic field (not the bird’s body) would rotate in the reference frame.
Magnetic field Bext, whose included angle θ with the x-axis changes from -180° to 180°, was
applied as a boundary condition to simulate the geomagnetic field. Assuming that magnetic
field Bext is in the xz plane, the x component of the external magnetic field is Bext_x=B0cosθ,
the y component is Bext_y=0, and the z-component is Bext_z=B0sinθ, where B0=Bext=50 μT.
With regard to the magnetic material properties, two kinds of magnetic materials were
proposed to be involved in the magnetoreception process.
a) SP nanoparticles: This proposal was based on the findings of the SP nanoparticles in birds
using transmission electron microscope (TEM) imaging in bright and dark-field mode and
using small-area electron diffraction (SAED) [4]. The SP particles, whose coercivity is close
to zero, were considered as an amplifier of
the geomagnetic field during
the
magnetoreception process. Considering the lack of measurement results of the specific
properties of magnetic particles in birds, the experimental results of the SP nanoparticle
(diameter: 10 nm, produced by Ferrotec Corporation, Japan) was used to determine the
magnetic properties. The B–H curve shown in Fig. 1(b) was measured using a vibrating
sample magnetometer (VSM). The coercivity of the SP material for the magnetic field
calculation obtained from the B–H curve is zero, and the relative permeability is 16. The
measurement sample of the SP nanoparticle was a magnetic fluid and not pure SP
nanoparticles. To obtain the magnetic properties of the SP materials, the curve was amplified
using the saturation magnetization of Fe3O4 (Ms=480 kA/m).
b) Ferromagnetic materials: This proposal was based on the findings of two different types of
iron compounds in the upper beak skin of adult homing pigeons using different light and
electron microscopic methods combined with X-ray analysis. These two iron compounds
were identified as two ferromagnetic materials: magnetite and maghemite [14], which closely
resemble the real iron-containing structure in birds. However, the specific magnetic
properties of the two types of materials have not been identified by measurements yet. In
comparison with the SP materials, the coercivity of the ferromagnetic materials for magnetic
field calculation was set to 17 kA/m, and the relative permeability was set to 16. In addition,
the residual magnetization direction of the magnetic cylinder was considered to be along the
x-axis (i.e., central axis of the cylinder). These parameters were almost the same as those used
in the magnetic particle simulation model of Solov’yov et al. [15].
The resultant magnetic fields corresponding to these two kinds of magnetic materials were
calculated and used to analyze the product yield of the radical pair. As this analysis was a
large-scale qualitative simulation, the assumption for the two magnetic materials did not
consider the special characteristics of the magnetic particles in the nanometer scale.
The isotropic radical-pair magnetic field effect was used to calculate the singlet state product
5
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
yield of the reactions under different magnetic field strengths using the method of Timmel et
al. [42]. This method was considered in the case of a radical pair with a single spin-1/2
nucleus (e.g., a proton), which ignored radical-pair diffusion and allowed singlet/triplet pairs
to disappear with the first-order kinetics with rate constants kS=kT=k [43].
The formula for calculating the singlet state product yield ΦS is shown in Eq. (1), where ω is
the Larmor frequency (ω=γB) and a is the isotropic hyperfine coupling constant.
2
⋅
⋅
a
1
3 1
2
ω
Φ = + ⋅
+ ⋅
s
8
8 8
2
2
Ω Ω
1
1
ω
f
)
(1
(
+ ⋅ −
2
8
Ω
1
1
ω
)
(1
+ ⋅ −
2
8
Ω
1
1
ω
)
(1
+ ⋅ +
8
2
Ω
1
1
ω
)
(1
+ ⋅ +
8
2
Ω
(
(
f
(
⋅
⋅
f
f
⋅
f
(
)
Ω
a
+
a
−
a
−
a
+
1
2
1
2
1
2
1
2
1
)
+
ω Ω
2
1
) (1)
−
ω Ω
2
1
)
+
ω Ω
2
1
)
−
ω Ω
2
In addition, f(x) is the Lorentzian function, where
f x
( )
=
k
2
aΩ
(
=
2
2
k
+
+
(2)
. (3)
2
x
)
2 1/ 2
ω
Ⅳ. CALCULATION RESULTS
A. Magnetic field distribution
The relationship among the amplitude of the resultant magnetic field B, the included angle θ
between the applied magnetic field and the central axis of the magnetic cylinder, and length
Li in curves l1 and l2, calculated in the cylindrical magnet of SP and the ferromagnet, is shown
on the 3-D diagram in Fig. 2. Li is defined as the length away from point ai in curve li (i=1, 2).
When θ is 0°, the B distribution corresponding to distance L1 in curve l1 is the same as that
of L2 in curve l2 because of axial symmetry, as shown in Fig. 3(a) and (b). The relationship
between B and dx (distance from o' to the calculation point at the x-axis) is shown in Fig.
3(c).
To investigate the effect of the change in the magnetic field amplitude (ΔB=Bmax-Bmin,
where Bmax is the maximum value of B and Bmin is the minimum value of B) on the
product yield of a radical pair, B of centerpoint c in curves l1 and l2 (i.e., Li≈800 μm) in the
SP and ferromagnetic materials was calculated. This process also compared the difference
between the two materials. The relationships between B and θ in the SP and ferromagnetic
materials with dx=0.4 mm are shown in Fig. 4(a) and (b), respectively (ΔBSP≈0.03 μT and
ΔBFerromagnet≈24 μT).
6
]
T
μ
[
B
65
60
55
50
45
40
35
1600
1200
800
400
180
120
60
0
-60
-120
0
-180
θ [°]
(a) (b)
L1 [μm]
0
-180
-120
180
120
60
0
-60
θ [°]
1200
800
400
L1 [μm]
]
T
μ
[
B
50.04
50.03
50.02
50.01
50.00
1600
]
T
μ
[
B
50.04
50.03
50.02
50.01
50.00
1600
65
55
45
]
T
μ
[
B
35
1600
1200
180
120
60
800
400
1200
800
400
L2 [μm]
0
-180
-120
0
-60
θ [°]
L2 [μm]
(c) (d)
0
-180
180
120
60
-120
0
-60
θ [°]
1
2
3
4
5
6
7
8
Fig 2. Resultant magnetic field B distribution when θ varies from -180° to 180°.
(a) Relationship among B, θ, and the position of the observed point in curve l1 in the SP material. (b)
Relationship among B, θ, and the position of the observed point in curve l1 in the ferromagnetic material. (c)
Relationship among B, θ, and the position of the observed point in curve l2 in the SP material. (d) Relationship
among B, θ, and the position of the observed point in curve l2 in the ferromagnetic material.
65
60
55
50
45
40
35
]
T
μ
[
B
SP
Ferromagnet
50.04
50.03
50.02
50.01
50.00
]
T
μ
[
B
0
400
800
L [μm]
1200
1600
SP
Ferromagnet
0
400
800
L [μm]
1200
1600
(a) (b)
7
SP
Ferromagnet
1000
800
600
400
200
]
T
μ
[
B
0
0
0.2
0.4
0.6
dx [mm]
(c)
0.8
1
Fig 3. Distribution of B at different positions when θ=0°.
(a) Distribution of B along curve l1 or l2. (b) Partial enlargement of Fig. 4(a). (c) Distribution of B along the
x-axis.
50.04
50.03
50.02
50.01
]
T
μ
[
B
SP
50.00
-180 -120
-60
60
120
180
0
θ [o]
]
T
μ
[
B
65
60
55
50
45
40
35
-180 -120
Ferromagnet
-60
0
θ[o]
60
120
180
(a) (b)
Fig 4. Relationship between B and θ at dx=0.4 mm.
(a) In the SP material. (b) In the ferromagnetic material.
B. Singlet state product yield corresponding to the magnetic field
The singlet state product yield ΦS was calculated according to the method mentioned in
Section III. The relationship between the singlet state product yield ΦS and B is shown in Fig.
5(a). To facilitate discussion, the horizontal axis considered in the figure was B and not B/a,
although the latter is more commonly used in other papers. In addition, based on the
experimental results in a protein environment using flash photolysis [44], the decay rate k
was set as 1 μs-1. According to the calculation formula, the different values of hyperfine
coupling constant a influenced the ΦS range corresponding to the change in the magnetic
field direction angle θ from -180° to 180° located in different parts of the curve, as shown in
Fig. 5(b) and (c). If a is much higher than the Larmor frequency of B, the ΦS range
corresponding to the magnetic field direction angle θ monotonically decreases, as shown by
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
8
80
Bmin Bmax
70
]
%
[
S
Φ
60
0
a=1.6μs-1
a=2.1μs-1
a=3.0μs-1
50
100
B [μT]
150
200
(a)
a=1.6μs-1
a=2.1μs-1
a=3.0μs-1
//
//
]
%
[
S
Φ
60
120
180
//
75.00
74.50
74.00
73.50
73.00
72.50
67.30
67.20
67.10
67.00
62.50
62.00
61.50
61.00
60.50
-180 -120
//
-60
a=1.6μs-1
a=2.1μs-1
a=3.0μs-1
//
//
60
120
180
0
θ [ο]
]
%
[
S
Φ
73.2030
73.2025
73.2020
73.2015
73.2010
//
67.0541
67.0540
//
61.4615
61.4610
61.4605
61.4600
61.4595
61.4590
-180 -120
-60
0
θ [o]
(b) (c)
Fig 5. Variation in the singlet state product yield ΦS of the radical pairs.
(a) Relationship between ΦS and B/a. (b) Relationship between ΦS and θ in the SP material. (c) Relationship
between ΦS and θ in the ferromagnetic material.
the dashed line (a=3.0 μs-1) in Fig. 5(a). The relationships between ΦS and θ in the SP and the
ferromagnetic materials are shown by the dashed lines in Fig. 5(b) and (c), respectively. If a
is much lower than the Larmor frequency of B, the ΦS range corresponding to the magnetic
field direction angle θ monotonically increases, as shown by the dot–dashed line (a=1.6 μs-1)
in Fig. 5(a). The relationships between ΦS and θ in the SP and ferromagnetic materials are
shown by the dot–dashed lines in Fig. 5(b) and (c), respectively. If the value of a is
appropriate, the ΦS range can also include the turning point, as shown by the solid line (a=2.1
μs-1) in Fig. 5(a). Under this condition, the relationship between ΦS and θ in the SP and
ferromagnetic materials are shown by the solid lines in Fig. 5(b) and (c), respectively.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
9
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
C. Modulation patterns corresponding to the geomagnetic information
The results presented earlier only indicated the variation of ΦS corresponding to the magnetic
field direction at a specific point. Generally, the sensory receptors of organisms are assumed
as ordered structures; thus, a reception plane is chosen to investigate the singlet state product
yield distribution of the ferromagnetic material in this section. Assuming that the residual
magnetization direction of the magnetic cylinder is along the bird’s body axis, the sensory
receptors are arranged in reception plane S (area: 0.5 mm×0.5 mm) parallel to the yz plane.
The distance of reception plane S along the positive direction of the x-axis from o' is 0.4 mm,
as shown by the gray quadrangle in Fig. 1(a). According to the proposed model, when the
applied magnetic field rotates parallel to the xy and xz planes (the included angle θ with the
x-axis changes from 0° to 180°), the ΦS distribution corresponding to B is calculated as
shown in Fig. 6.
67.05
67.10
67.15
67.20
67.25
67.30 [%]
0°
30°
60°
90° 120° 150° 180°
xy
xz
Fig. 6. Modulation patterns of ΦS through the magnetic field parallel to the xy and xz planes in different
directions at angles 0, 30, 60, 90, 120, 150, and 180°.
Ⅴ. DISCUSSION
Compared with the research conducted by Cohen and Cai, the difference in the combined
mechanisms in this paper is that we focused on the amplitude of the resultant magnetic field
around the magnetic particles instead of the magnetic field gradient. This process means that
the changes in the singlet state product yields correspond directly to the magnetic field at the
same spatial position, and the difference in the local magnetic fields between the two radicals
of each radical pair was ignored. This result can be attributed to the difference in the local
magnetic fields of the two radicals, which is much smaller than the changes in the resultant
magnetic field amplitude ΔB during the geomagnetic field rotation in the proposed
mechanism. For example, when the distance from the magnetic particles dx=0.4 mm,
ΔBSP≈0.03 μT, and ΔBFerromagnet≈24 μT. If the distance between two radicals is 3.5 nm (the
same as the value used in Cai’s calculation [39]), the variation in the local magnetic fields of
10
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
the two radicals is approximately 5.29×10-5 μT. In Cai’s calculation, it was approximately 4
mT, considerably larger than our calculation result, possibly because Cai assumed that the
acceptor of the radical pairs was 35 nm away from the magnetic particles. Therefore, the
proposed model appears more reasonable in cases when the distance between the radical pairs
and the magnetic particles is relatively farther.
Based on the magnetic field distribution calculation results, B was found to change
cyclically with included angle θ, and the period of variation was 180° in the SP materials and
360° in the ferromagnetic materials. When θ was 0°, B of centerpoint c in the two kinds of
materials both reached the maximum. The BSP variation was similar to the function
f(θ)=sin(θ−90°)+offset. The BFerromagnet variation at centerpoint c was similar to cosine
waves with an offset because ferromagnetic materials have inherent remanence, whereas SP
materials do not have any.
In the calculation, ΔBSP is very small because B decays exponentially with dx; if dx is
smaller, then ΔBSP would become higher. For example, if dx=0.1 mm, then ΔBSP≈0.3 μT,
and if dx=10 μm, then ΔBSP≈20 μT. In all cases, the variation trends of B in the SP material
with different dx’s are almost similar. Therefore, the calculation result with dx=0.4 mm can
still be used for qualitative analysis.
At the same spatial position as in point c in Fig. 1(a), BFerromagnet is much larger than BSP.
The change in the singlet state product yields ΔΦS = ΦSmax- ΦSmin of the ferromagnetic
material is also much larger than ΦS of the SP material. When dx=0.4 mm (ΔBSP≈0.03 μT
and ΔBFerromagnet≈24 μT), the decay rate is k=1 μs-1, and the values of hyperfine coupling
constant a are 1.6, 2.1, and 3.0 μs-1. The calculated ΔΦS of the SP materials are
approximately 1.8×10-5, 3.6×10-7, and 2.4×10-5, and ΔΦS of the ferromagnetic materials are
approximately 1.36%, 0.24%, and 1.91%, respectively. The changes in product yields ΦS
calculated at present are relatively small. If dx decreases slightly (but not below micrometer
level), then ΔΦS would become more significant. For example, when dx=0.1 mm, ΔΦS of the
SP material corresponding to the same values of hyperfine coupling constant a of 1.6, 2.1,
and 3.0 μs-1 would be approximately 1.8×10-4, 1.2×10-6, and 2.4×10-4, respectively
(ΔBSP≈0.3 μT). ΔΦS of the ferromagnetic materials are approximately 6.56%, 3.03%, and
2.59%, respectively (ΔBFerromagnet≈250 μT), which would mean that the reception radical
pairs are probably closer to the magnetic particles and not located in the retina of the birds,
according to the proposed mechanism.
We can reasonably assume that nature has optimized the properties through evolution to
provide maximum results; therefore, these calculation results can explain why biophysical
experiments showed that the magnetic particles in organisms are more similar to
ferromagnetic (and not SP) materials [14]. The former can provide more significant changes
in product yield ΦS, which made magnetoreception more feasible.
Moreover, when the range of singlet state product yield ΦS includes a turning point in the
ferromagnetic material or the ΦS range monotonically decrease or increase in the SP material,
the values of ΦS corresponding to the magnetic field direction angle θ=0° and θ=180° both
attained the maxima (the values are almost equal), indicating that the magnetic compass is an
inclination compass. This conclusion is in agreement with the experimental results obtained
from a variety of research activities on migratory birds [45], where the formula for
calculating the singlet state product yield ΦS was based on one-proton radical pair. In fact,
11
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
when the radical pairs are more complex, the calculation formula will also become more
complex. However, the curve of ΦS corresponding to B also has an extreme value; therefore,
the proposed model is still suitable.
In addition, because the experimental results showed that the properties of the magnetite
crystals in birds are more similar to ferromagnetic materials, SD magnetic materials were not
considered in the model calculation. In fact, SD magnetic materials will only make the
magnetic field around the magnetic particles higher and, hence, induce more significant
changes in the product yield, but the variation trend of ΦS calculated with SD materials is still
similar to that of the ferromagnetic materials.
Based on the response pattern of the singlet state product yield ΦS, the geomagnetic
information was found to be represented by the distribution of ΦS in the reception plane. The
position of high ΦS is located at the center position of the reception plane because the body
axis is parallel to the geomagnetic field. If the included angle between the body axis and the
geomagnetic field is changed in the horizontal or vertical plane, the position of the high ΦS
will move correspondingly along the horizontal or vertical direction. The pattern for a bird
flying anti-parallel to the magnetic field direction (180°) is similar to the pattern for parallel
orientation (0°), which also demonstrates that the magnetic compass of the birds is
intrinsically an inclination compass. Birds can use modulation patterns to adjust the flight
direction.
The previous discussions ignored the rotation of the bird itself around the body axis during
the flight, which often happens during the bird’s attitude adjustment process. However, the
bird can receive the rotation information from multiple magnetic particle arrangements or
from other means (such as the gravity information response) to assist in navigation.
Furthermore, the modulation patterns are very similar to the results obtained by
radical-pair-based mechanism only [7, 8], but the biophysical mechanisms are completely
different.
Ⅵ. CONCLUSION
This paper has proposed a biophysical mechanism for magnetoreception in birds, which
combined the magnetite- and radical-pair-based mechanisms to provide a more reasonable
explanation of avian magnetoreception. The amplitude of the resultant magnetic field
produced by the interaction between the magnetic particles and the geomagnetic field was
analyzed using two materials with different magnetic properties. The relationship between the
direction of the geomagnetic field and the radical-pair product yield was investigated. The
calculation results showed that the period of variation for singlet state products is
approximately 180° when the magnetic particles are ferromagnetic materials, which could
explain the experimental results of avian magnetic compass research. The singlet state
products through the magnetic field in a reception plane were investigated. Their modulation
patterns can provide geomagnetic information in birds. The model calculation results showed
that the proposed model, which deals with the magnetic particles’ role in changing the
direction information of the geomagnetic field to amplitude of the resultant magnetic field,
can explain how birds use geomagnetic field for navigation better than using either
magnetite- or radical-pair-based mechanism only.
12
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
ACKNOWLEDGMENTS
The authors thank Dr. Chunxiao Xu for valuable discussions. This work was supported by the
State Key Program of National Natural Science of China (51037006). This work was
supported by the State Key Development Program for Basic Research of China
(2011CB503702).
References
[1] W. Wiltschko, and F. Merkel, Verh. dt. zool. Ges. 59 (1966).
[2] S. Johnsen, and K. J. Lohmann, Nat Rev Neurosci 6 (2005).
[3] J. L. Kirschvink et al., Bioelectromagnetics (1992).
[4] H. Cadiou, and P. A. McNaughton, J R Soc Interface 7 (2010).
[5] V. P. Shcherbakov, and M. Winklhofer, Phys Rev E 81 (2010).
[6] K. Schulten, C. E. Swenberg, and A. Weller, Z Phys Chem Neue Fol 111 (1978).
[7] T. Ritz, S. Adem, and K. Schulten, Biophys J 78 (2000).
I. A. Solov'yov, H. Mouritsen, and K. Schulten, Biophys J 99 (2010).
[8]
[9] R. Blakemore, Science 190 (1975).
[10] M. M. Walker et al., Nature 390 (1997).
[11] J. L. Kirschvink, Nature 390 (1997).
[12] J. L. Kirschvink, M. M. Walker, and C. E. Diebel, Curr Opin Neurobiol 11 (2001).
[13] A. F. Davila et al., Phys Chem Earth 28 (2003).
[14] G. Fleissner et al., Naturwissenschaften 94 (2007).
[15] I. A. Solov'yov, and W. Greiner, Biophys J 93 (2007).
[16] R. C. Beason, and P. Semm, Neurosci Lett 80 (1987).
[17] R. C. Beason, and P. Semm, J Exp Biol 199 (1996).
[18] M. J. M. Leask, Nature 267 (1977).
[19] R. Wiltschko, and W. Wiltschko, Naturwissenschaften 85 (1998).
[20] W. Wiltschko et al., Nature 364 (1993).
[21] M. Zapka et al., Nature 461 (2009).
[22] K. M. Salikhov et al., Elsevier (1984).
[23] U. E. Steiner, and T. Ulrich, Chem Rev 89 (1989).
[24] C. R. Timmel et al., Mol Phys 95 (1998).
[25] C. R. Timmel et al., Chem Phys Lett 334 (2001).
[26] C. T. Rodgers, and P. J. Hore, P Natl Acad Sci USA 106 (2009).
[27] T. Ritz et al., J R Soc Interface 7 (2010).
[28] J. C. S. Lau et al., J R Soc Interface 7 (2010).
[29] R. P. Groff et al., Phys Rev B 9 (1974).
[30] W. Bube et al., Chem Phys Lett 50 (1977).
[31] S. G. Boxer, C. E. D. Chidsey, and M. G. Roelofs, P Natl Acad Sci-Biol 79 (1982).
[32] B. Vandijk, R. Vandervos, and A. J. Hoff, Chem Phys Lett 226 (1994).
[33] K. Maeda et al., Nature 453 (2008).
[34] C. Niessner et al., Plos One 6 (2011).
[35] V. N. Binhi, Bioelectromagnetics 27 (2006).
[36] V. N. Binhi, Int J Radiat Biol 84 (2008).
[37] A. E. Cohen, J Phys Chem A 113 (2009).
13
1
2
3
4
5
6
7
8
9
10
11
[38] N. Yang, and A. E. Cohen, Opt Express 18 (2010).
[39] J. M. Cai, G. G. Guerreschi, and H. J. Briegel, Phys Rev Lett 104 (2010).
[40] M. Winklhofer, and J. L. Kirschvink, J R Soc Interface 7 (2010).
[41] G. Falkenberg et al., Plos One 5 (2010).
[42] C. R. Timmel, and K. B. Henbest, Philos T Roy Soc A 362 (2004).
[43] A. J. Hoff, and P. J. Hore, Chem Phys Lett 108 (1984).
[44] N. Mohtat et al., Photochem Photobiol 67 (1998).
[45] W. Wiltschko, and R. Wiltschko, J Exp Biol 199 (1996).
14
|
1212.2206 | 2 | 1212 | 2013-01-02T14:40:17 | Structure-property-function relationships in triple helical collagen hydrogels | [
"physics.bio-ph",
"cond-mat.mtrl-sci",
"cond-mat.soft",
"physics.chem-ph",
"q-bio.BM",
"q-bio.TO"
] | In order to establish defined biomimetic systems, type I collagen was functionalised with 1,3-Phenylenediacetic acid (Ph) as aromatic, bifunctional segment. Following investigation on molecular organization and macroscopic properties, material functionalities, i.e. degradability and bioactivity, were addressed, aiming at elucidating the potential of this collagen system as mineralization template. Functionalised collagen hydrogels demonstrated a preserved triple helix conformation. Decreased swelling ratio and increased thermo-mechanical properties were observed in comparison to state-of-the-art carbodiimide (EDC)-crosslinked collagen controls. Ph-crosslinked samples displayed no optical damage and only a slight mass decrease (~ 4 wt.-%) following 1-week incubation in simulated body fluid (SBF), while nearly 50 wt.-% degradation was observed in EDC-crosslinked collagen. SEM/EDS revealed amorphous mineral deposition, whereby increased calcium phosphate ratio was suggested in hydrogels with increased Ph content. This investigation provides valuable insights for the synthesis of triple helical collagen materials with enhanced macroscopic properties and controlled degradation. In light of these features, this system will be applied for the design of tissue-like scaffolds for mineralized tissue formation. | physics.bio-ph | physics | Structure-property-function relationships in triple-helical collagen hydrogels
Giuseppe Tronci,1,2 Amanda Doyle,1,2 Stephen J. Russell,2 and David J. Wood1
1Biomaterials and T issue Engineering Research Group, Leeds Dental Institute, University o f
Leeds, Leeds LS2 9LU, United Kingdom
2 Nonwoven Research Group, Centre for Technical Textiles, University o f Leeds, Leeds LS2
9JT, United Kingdom
Abstract
In order to establish defined biomimet ic systems, type I collagen was funct ionalised with
1,3-Phenylenediacet ic acid (Ph) as aromat ic, bifunct ional segment. Following invest igat ion on
mo lecular organizat ion and macroscopic properties, material functionalit ies, i.e. degradability
and bioact ivity, were addressed, aiming at elucidating the potential o f this co llagen system as
mineralizat ion template. Funct ionalised co llagen hydrogels demonstrated a preserved triple helix
conformation. Decreased swelling rat io and increased thermo -mechanical properties were
observed in comparison to state-of-the-art carbodiimide (EDC)-crosslinked co llagen controls.
Ph-crosslinked samples displayed no optical damage and only a slight mass decrease (~ 4 wt. -%)
fo llowing 1-week incubat ion in simulated body fluid (SBF), while nearly 50 wt. -% degradation
was observed in EDC-crosslinked collagen. SEM/EDS revealed amorphous mineral deposit ion,
whereby increased calcium phosphate ratio was suggested in hydrogels with increased Ph
content. This invest igat ion provides valuable insights for the synthesis of triple helical co llagen
materials with enhanced macroscopic properties and controlled degradat ion. In light of these
features, this system will be applied for the design o f tissue -like scaffo lds for mineralized t issue
format ion.
1. Introduction
Collagen is the main protein of the human body, ruling structure, function and shape of
bio logical t issues. Also in light of it s unique mo lecular organization, co llagen has been widely
applied for the design o f vascular grafts [1], fibrous materials for stem cell different iat ion [ 2],
biomimet ic scaffo lds for regenerat ive medicine [3], and t issue-like matrices for hard t issue repair
[4]. However, collagen properties are challenging to control in phys io logical condit ions, mainly
because its hierarchical organization and chemical composit ion in vivo can only be part ially
reproduced in vitro. Funct ionalisat ion and crosslinking of collagen mo lecules, e.g. via
carbodiimide [5,6], glutaraldehyde [7,8] or hexamethylene diisocyanate [9], have proved to
enhance macroscopic properties in aqueous environment, although much is st ill left to do to
establish biomimet ic systems with defined structure-property-funct ion relat ionships. Here, the
design of type I collagen hydrogels was investigated via covalent lysine funct ionalisat ion with
1,3-Phenylenediacet ic acid (Ph). It was hypothesized that incorporation of a st iff, aromat ic
segment among co llagen mo lecules could offer a novel synthetic route to the formation o f
mechanically-relevant materials. Ph was selected as bifunct ional segment, in order to promote
crosslinking o f distant collagen mo lecules, unlikely accomplished with current synthetic methods
[1,5,6], so that triple helix conformat ion of collagen could be retained. The pr esence o f an
aromat ic ring in the Ph was considered crucial to achieve controlled swelling and enhanced
mechanical properties in result ing hydrogels, owing to the mo lecular st iffness o f incorporated
segment. In order to invest igate the effect iveness of this synthetic approach, EDC treatment was
selected as state-of-the-art reference method, since it has been shown to promote the format ion
of water-stable co llagen materials with no residual toxicity, in contrast to aldehyde biomateria l
fixat ion [1]. Formed hydrogels were invest igated for mo lecular conformat ion, network
architecture and hydrogel macroscopic properties, in comparison to EDC-crosslinked co llagen.
Furthermore, material degradability and bioact ivity were addressed via incubat ion in SBF, in
order to elucidate materials’ potential for biomime tic mineralizat ion.
2. Experimental details
1,3-Phenylenediacet ic acid (Ph) was supplied by VWR Internat ional, all the other chemicals
were purchased from Sigma-Aldrich. Type I collagen was iso lated in-house from rat tail tendons
[10]. Collagen was disso lved in 10 mM hydrochloric acid and funct ionalised with N -(3-
Dimethylaminopropyl)-N′-ethylcarbodiimide hydrochloride (EDC)-act ivated Ph under gentle
shaking at room temperature. EDC-crosslinked collagen was synthesized by mixing EDC with
collagen so lution, as previously reported [5]. Result ing hydrogels were washed with dist illed
water and dehydrated in aqueous solut ions o f increasing ethano l concentrations. Attenuated Total
Reflectance Fourier Transform Infrared (ATR-FTIR) spectroscopy was carried out on dry
samples using a Perkin-E lmer Spectrum BX spotlight spectrophotometer with diamond ATR
attachment. 64 scans were averaged for each spectrum, using 4 cm−1 reso lut ion and 2 cm−1
scanning interval. Degree o f crosslinking (C) of co llagen networks was determined by
2,4,6-trinitrobenzenesulfonic acid (TNBS) co lorimetric assay (n=2) [11], as the mo lar ratio
between funct ionalised and pristine, non functionalised lys ines. Swelling tests (n=3) were carried
out by incubat ing dry samples in 5 mL dist illed water for 24 hours. Water -equilibrated samples
were retrieved, paper-blotted and weighed. The weight -based swelling rat io (SR) was calculated
as SR=(ms-md)/md ·100, where ms and md are swollen and dry sample weights, respect ively.
Hydrogel discs (ø 0.8 cm, n=4) were compressed ( Instron 5544 UTM) with a compression rate of
3 mm·min-1. Different ial Scanning Calorimetry (DSC) temperature scans were conducted on 10 -
140 °C temperature range with 10 °C·min-1 heat ing rate (TA Instruments Thermal Analysis 2000
System and 910 Different ial Scanning Calorimeter cell base). Mineralization experiment was
carried out via 1-week sample incubat ion (n=4) at 25 °C SBF, with 0.7 sample weight/solut ion
vo lume ratio [12]. Retrieved samples were rinsed with dist illed water, dried, weighed and gold -
coated for SEM/EDS (JEOL SM-35) analys is.
3. Discussion
3.1 Molecular organization of crosslinked collagen via ATR-FTIR spectroscopy
Triple helix co llagen conformation is normally associated with three main amide bands, i.e.
amide I at 1650 cm−1, result ing from the stretching vibrat ions o f peptide C=O groups; amide II
absorbance at 1550 cm−1, deriving from N–H bending and C–N stretching vibrations; and amide
III band centered at 1240 cm−1, assigned to the C–N stretching and N–H bending vibrat ions from
amide linkages, as well as wagging vibrat ions of CH2 groups in the glycine backbone and proline
side chains [13]. Figure 1 indicates that the posit ions of these amide bands are maintained in Ph-
as well as EDC-crosslinked networks. Furthermore, the FTIR absorption ratio of amide III to
1450 cm−1 band was determined to be close to unity (AIII/A1450 ~ 1.01-1.14) among the three
samples, suggest ing preserved integrity o f triple helices [14] fo llowing funct ionalisat ion of
nat ive co llagen.
Figure 1. Exemplary FTIR spectra of Ph- (black line, left) and EDC-crosslinked (blue line, right) collagen. Native
collagen (light gray line) spectrum is displayed for comparison in both plots.
3.2 Network architecture and macroscopic properties of collagen hydrogels
Network architecture was invest igated both by quant ifying the degree of crosslinking (C) via
TNBS assay and by assessing the swelling rat io of result ing materials. As observed in Table I,
collagen-Ph samples displayed a higher degree of crosslinking compared to state-of-the-art
EDC-crosslinked co llagen, despite having a very low Ph-collagen lysine mo lar rat io. A slight
increase of Ph content (0.5 to 1.5 [COOH]/[Lys] ratio) in the cro sslinking mixture led to
nearly-complete functionalisat ion of co llagen lys ines (> 99 mo l.-%). These results suggest that
the employment of a bifunct ional segment is likely to promote an increased yield o f co llagen
funct ionalisat ion, since co llagen mo lecules separated by a distance can be bridged [8]. On the
other hand, EDC-mediated functionalisat ion results in the format ion o f zero-length net-points,
whereby intramo lecular crosslinks are likely established [5,8]. Consequently, steric hindrance
effects almo st certainly explain the decreased degree of crosslinking observed in co llagen-EDC,
compared to collagen-Ph, samples. Besides the degree o f crosslinking, hydrogel swelling
behaviour was also investigated; the swelling rat io of one sample, co llagen-Ph1.5, was found to
be significant ly lower compared to all other EDC-based samples. This is supported by TNBS
results, suggest ing nearly complete functionalisat ion of co llagen lysines for this composit ion. As
for the other Ph-crosslinked samples (co llagen-Ph0.5/1), swelling rat ios were similar to each
other, which is again supported by TNBS results, describing a similar degree of crosslinking.
In order to invest igate whether variation o f mo lecular parameters could induce changes in
macroscopic properties, the thermo -mechanical properties of co llagen hydrogels were addressed.
Table I. Degree of crosslinking (C), swelling ratio (SR) and denaturation temperature (Td) of collagen hydrogels.
*Sample are coded as ‘Collagen -XXX-YY’, where XXX indicates the type of system (either Ph or EDC-based),
while YY identifies the molar ratio of either Ph carboxylic functions or EDC to collagen lysines.
Sample ID*
C /mol.-% SR /wt.-% Td /°C
Collagen-Ph0.5
88 ± 3
1285 ± 450
Collagen-Ph1
87 ± 14
1311 ± 757
Collagen-Ph1.5
> 99
823 ± 140
Collagen-EDC10
25 ± 6
1392 ± 82
Collagen-EDC20
37 ± 13
1595 ± 374
Collagen-EDC30
34 ± 1
1373 ± 81
Collagen-EDC40
68 ± 3
1374 ± 182
Collagen-EDC60
60 ± 1
1106 ± 82
80
80
88
68
76
78
80
80
Collagen denaturation temperature (Td) is related to the unfo lding o f co llagen triple helices into
randomly-co iled chains; it is therefore expected to be highly affected by the format ion o f a
covalent network [2]. Table I describes hydrogel Td values as obtained by DSC; crosslinked
samples show a denaturation temperature in the range o f 68–88 °C, which is found to be higher
than the denaturation temperature of native co llagen (Td ~ 67 °C). Furthermore, variat ion o f Td
seems to be direct ly related to changes o f crosslinking degree in the hydrogel network. These
results give supporting evidence that covalent net -points were established during hydrogel
format ion, so that collagen triple helices were successfully retained and stabilized. It should be
noted that Ph-crosslinked co llagen revealed higher denaturation temperatures with respect to
EDC- [6], glutaraldehyde- [7], and hexamethylene diisocyanate- [9] crosslinked co llagen
materials. This indicates that the incorporation of Ph as a st iff, aromat ic segment superiorly
stabilizes co llagen mo lecules in comparison with current crosslinking methods.
Besides thermal analys is, mechanical properties o f co llagen-Ph hydrogels were measured by
compression tests. Samples described J-shaped stress-compression curves (data not shown),
similar to the case o f nat ive tissues. Here, shape recovery was observed fo llowing load remova l
up to nearly 50% compression, suggest ing that the established covalent network successfully
resulted in the format ion of an elast ic material, as observed in linear biopo lymer networks [15].
On the other hand, EDC-crosslinked co llagen showed minimal mechanical properties, whereby
sample break was observed even after sample punching. Consequent ly, quant itative data of
mechanical properties on co llagen-EDC samples could not be acquired. These findings give
further evidence that the collagen funct ionalisation with Ph is effect ive for the format ion o f
collagen materials with enhanced mechanical properties. Compressive modulus
(E:
28±1035±9 kPa) and maximal stress of Ph-crosslinked samples (σmax: 6±28±4 kPa) were
measured in the kPa range, while compression at break (εb: 53±5–58±5 %) did not exceed 60%
compression (Figure 2). The resulting compressive modulus was therefore measured to be almost
20 times higher compared to previously-reported collagen-based materials [8].
Figure 2. Compressive modulus (E), maximal compressive stress (σmax), and compression at break (εb) of
Ph-crosslinked collagen hydrogels (n=4).
At the same t ime, there was litt le variat ion in mechanical properties among the different
composit ions. Given the unique organization o f collagen, the hierarchical level at which covalent
crosslinks are introduced is crucial in order to study the influence o f crosslinking on the
mechanical properties o f co llagen. Olde Damink et al. observed no variat ion o f mechanica l
properties in dermal sheep crosslinked co llagen [6,7,9]. This was explained based on the fact that
crosslinks were mainly introduced within rather than among co llagen mo lecules. This hypothesis
may be supported by above mechanical findings, although it is not in line with TNBS, swelling
and thermal analys is data. Most likely, the variat ion of Ph feed ratio among the different
composit ions (0.51.5 [COOH]/[Lys] ratio) was probably too low to result in significant
changes in mechanical properties. For these reasons, a wider range of Ph concentrations may be
advantageous in order to establish hydrogels with varied mechanical properties. In that case,
investigat ion via AFM will be crucial in order to explore the hierarchical levels at which
covalent crosslinks are introduced.
3.3 SBF incubation of collagen hydrogels
SBF incubation is a well-known method to test a material’s ability to form a
hydroxycarbonate apatite (HCA) layer in vitro [12]. Collagen is known to trigger bone-like
apatite deposit ion in vivo during bone format ion [4], so it was of interest to invest igate collagen
hydrogel behaviour in SBF as an osteogenic-like medium. The mass change as well as the
presence o f calcium/phosphorous elements was therefore quant ified in retrieved samples in order
to (i) clarify any occurrence of degradation and (ii) determine the chemical composit ion of any
potentially-nucleated phases.
A slight decrease in mass (averaged mass loss ~ 4 wt.-%) was observed in co llagen-Ph
samples, indicat ing minima l hydro lyt ic degradation had occurred. SEM on retrieved samples
revealed nearly- intact material surfaces, confirming that a covalent network was still present at
the mo lecular level. Among the different composit ions, only one sample, collagen -Ph1,
displayed a slight mass increase; observed differences in hydro lyt ic degradat ion may be related
to a varied crosslinking/graft ing ratio in the formed hydrogel networks. Networks with increased
yield o f graft ing will likely degrade faster compared to networks with increased yield o f
crosslinking, since grafted mo lecules are expected to be cleaved more easily by water compared
to crosslinked mo lecules. Consequently, variat ion o f Ph feed rat io may affect the yield o f
crosslinking, so that grafted as well as crosslinked molecules may be formed above a specific Ph
ratio threshold. A much higher mass loss (averaged mass loss ~ 53 wt. -%) was observed in EDC-
compared to Ph-crosslinked co llagen, which is in line with previous findings. Here, small micro -
pores were observed on the retrieved sample surface, suggest ing a surface, rather than bulk,
erosion mechanism of hydro lyt ic degradat ion.
Besides degradation behaviour, SEM/EDS analyses were carried out to explore whether any
mineral phase was nucleated fo llowing SBF incubat ion. Here, sample washing with w ater was
crucial in order to remove superfic ial deposit ion of magnesium, sodium and chlorine ions, in
agreement with the use of SBF. The presence of calcium and phosphorous elements was
observed in all samples, although with a low Ca/P atomic ratio (0.84 -1.41). This suggests that the
mineral phase laid down on the material surface was most likely const ituted of amorphous
calcium phosphate, which may be expected due to the relat ively short incubation t ime (1 week)
at room instead of body temperature. Interestingly, sample collagen-Ph1.5 displayed increased
Ca/P atomic rat io (Ca/P ~ 1.41) compared to the other samples, likely hint ing at enhanced and
select ive nucleat ion of an apat ite layer in hydrogels crosslinked with increased Ph feed ratio.
4. Conclusions
This study highlights the important role played by network mo lecular architecture on the
macroscopic properties and funct ions of result ing collagen hydrogels. Fo llowing a bottom-up
synthetic approach, funct ionalisat ion with a bifunctional, aromat ic segment, successfully led to
the establishment of a biomimet ic system with preserved triple helix integrity and enhanced
macroscopic properties, in comparison to state-of-the-art crosslinked co llagen. Result ing
hydrogels displayed minimal hydro lyt ic degradation fo llowing 1-week incubat ion in SBF,
whereby nucleat ion of amorphous calcium phosphate phase was init iated.
Acknowledgements
This work was funded through WELMEC, a Centre of Excellence in Medica l
Engineering funded by
the Wellcome Trust and EPSRC, under grant number WT
088908/Z/09/Z. The authors would like to thank J. Hudson, W. Vickers and S. Finlay, for kind
assistance with SEM/EDS, SBF preparation, and compression tests, respectively.
References
1 T. Huynh, G. Abraham, J. Murray, K. Brockbank, P.-O. Hagen , S. Sullivan, Nature Biotechnology 17, 1084
(1999).
2 L. Meng, O. Arnoult, M. Smith and Gary E. Wnek, J. Mater. Chem. 22, 19414 (2012).
3 N. Davidenko, T. Gibb, C. Schuster, S.M. Best, J.J. Campbell, C.J. Watson, R.E. Cameron, Acta Biomater. 8, 667
(2012).
4 Y. Wang, T. Azaïs, M. Robin, A. Vallée, C. Catania, P. Legriel, G. Pehau -Arnaudet, F. Babonneau, M.-M. Giraud-
Guille, and N. Nassif, Nature Mater. 11, 724 (2012).
5 S. Yunoki and T. Matsuda, Biomacromolecules 9, 880 (2008).
6 L.H.H. Olde Damink, P.J. Dijkstra, M.J.A. van Luyn, P.B. van Wachem, P. Nieuwenhuis and J. Feijen,
Biomaterials 17, 772 (1996).
7 L.H.H. Olde Damink, P.J. Dijkstra, M.J.A. Van Luyn, P.B. Van Wachem, P. Nieuwenhuis and J. Feijen , J. Mater.
Sci. Mater. Med. 6, 465 (1995).
8 M.G. Haugh, C.M. Murphy, R.C. McKiernan, C. Altenbuchner, and F.J. O’Brien, Tissue Eng A Part A 17, 1202
(2011).
9 L.H.H. Olde Damink, P.J. Dijkstra, M.J.A. Van Luyn, P.B. Van Wachem, P. Nieuwenhuis and J. Feijen, J. Mater.
Sci. Mater. Med. 6, 431 (1995).
10 E. Bell, B. Ivarsson, and C. Merrill, Proc. Natl. Acad. Sci. 76, 1274 (1979).
11 W.A. Bubnis and C.M. Ofner, Analyt. Biochem. 207, 129 (1992).
12 S.K. Misra, T. Ansari, D. Mohn, S.P. Valappil, T.J. Brunner, W.J. Stark, I. Roy, J.C. Knowles, P.D. Sibbons, E.V.
Jones, A.R. Boccaccini and V. Salih, J. R. Soc. Interface 7, 454 (2010).
13 B.B. Doyle, E.G. Bendit, and E.R. Blout, Biopolymers 14, 940-944 (1975).
14 L. He, C. Mu, J. Shi, Q. Zhang, B. Shi, W. Lin, Int. J. Biol. Macromol. 48, 356 (2011).
15 G. Tronci, A.T. Neffe, B.F. Pierce, A. Lendlein, J. Mater. Chem. 20, 8881 (2010).
|
1811.10127 | 2 | 1811 | 2019-03-02T05:17:06 | An Explicit Electron-Vibron Model for Olfactory Inelastic Electron Transfer Spectroscopy | [
"physics.bio-ph"
] | The vibrational theory of olfaction was posited to explain subtle effects in the sense of smell inexplicable by models in which molecular structure alone determines an odorant's smell. Amazingly, behavioral and neurophysiological evidence suggests that humans and some insects can be trained to distinguish isotopologue molecules which are related by the substitution of isotopes for certain atoms, such as a hydrogen-to-deuterium substitution. How is it possible to smell a neutron? The physics of olfaction may explain this isotopomer effect. Inelastic electron transfer spectroscopy (IETS) has been proposed as a candidate mechanism for such subtle olfactory effects: the vibrational spectrum of an appropriately-quantized odorant molecule may enhance a transfer rate in a discriminating electron transfer (ET) process. In contrast to other semi-classical or quantum-master-equation-based models of olfactory IETS, the model presented here explicitly treats the dynamics of a dominant odorant vibrational mode, which provides an indirect dissipative path from the electron to the thermal environment. A direct dissipative path to the environment also is included. Within this model, a calculation of ET rate is developed, along with a calculation of power dissipation to the thermal environment. Under very weak direct dissipative coupling, spectroscopic behaviors of the indirect path are revealed, and the resulting ET rate exhibits resonant peaks at certain odorant frequencies. Resonant peaks in ET rate also correlate to peaks in power dissipation. Spectroscopic behaviors are masked by strong direct dissipative coupling. Results support a rate-based discrimination between a preferred ligand and an isotopomer if indirect dissipative coupling dominates. | physics.bio-ph | physics | An Explicit Electron-Vibron Model for Olfactory Inelastic Electron Transfer
Spectroscopy
Nishattasnim Liza1 and Enrique P. Blair1, a)
Electrical and Computer Engineering Department, Baylor University, One Bear Place #97356, Waco, Texas 76798,
USA
The vibrational theory of olfaction was posited to explain subtle effects in the sense of smell inexplicable
by models in which molecular structure alone determines an odorants smell. Amazingly, behavioral and
neurophysiological evidence suggests that humans and some insects can be trained to distinguish isotopologue
molecules which are related by the substitution of isotopes for certain atoms, such as a hydrogen-to-deuterium
substitution. How is it possible to smell a neutron? The physics of olfaction may explain this isotopomer effect.
Inelastic electron transfer spectroscopy (IETS) has been proposed as a candidate mechanism for such subtle
olfactory effects: the vibrational spectrum of an appropriately-quantized odorant molecule may enhance a
transfer rate in a discriminating electron transfer (ET) process. In contrast to other semi-classical or quantum-
master-equation-based models of olfactory IETS, the model presented here explicitly treats the dynamics of
a dominant odorant vibrational mode, which provides an indirect dissipative path from the electron to the
thermal environment. A direct dissipative path to the environment also is included. Within this model, a
calculation of ET rate is developed, along with a calculation of power dissipation to the thermal environment.
Under very weak direct dissipative coupling, spectroscopic behaviors of the indirect path are revealed, and
the resulting ET rate exhibits resonant peaks at certain odorant frequencies. Resonant peaks in ET rate
also correlate to peaks in power dissipation. Spectroscopic behaviors are masked by strong direct dissipative
coupling. Results support a rate-based discrimination between a preferred ligand and an isotopomer if indirect
dissipative coupling dominates.
I.
INTRODUCTION
Much insight into the mechanisms of other senses
has been developed, but mysteries persist in our mod-
ern understanding of olfaction.
It is known that odor-
ant molecules interact with olfactory receptors on sen-
sory neurons, and that an odorant's structure is im-
portant in determining its scent, but there is evidence
that additional, presently-unknown information may be
required to distinguish molecules. For example, isotop-
tomers -- molecules related by interchanging some atoms
with isotopes -- are nearly identical, yet some behav-
ioral experiments suggest that humans and fruit flies
can distinguish between isotopotomers by smell.1 -- 6 Still,
in vitro work also
the dispute remains unresolved:
performed in selected mammalian olfactory receptors
has not demonstrated a distinguishable response to iso-
topomer molecules,4 but neurophysiological research has
shown that insect antennae can produce a differential re-
sponse to some isotopomers.7,8
Olfactory models may be divided into two broad cat-
egories. One category may be termed "lock-and-key"
models, or "odotope" models, in which odorant structure
alone is assumed to determine the receptor response, just
as a key's shape determines whether it can actuate a lock.
A second class of models known as "swipe-card" models
extend the lock-and-key model in that structure is recog-
nized as necessary but insufficient to distinguish odors.9
One prominent swipe-card theory invokes quantum me-
chanics: a suitable structure and additional information
a)enrique [email protected]
encoded in the molecule trigger the olfactory response.
This is in analogy to a magnetic hotel key card, which
must not only fit in the swipe-card slot, but also must
have the proper keyword encoded in the magnetic strip.
Turin posited that spectroscopic information about a
molecule could enhance an electron transfer (ET) rate,
which would play an integral role in distinguishing be-
tween odorants.10,11 After entering the nasal cavity and
diffusing through the mucus layer, an odorant molecule
may dock with an odorant receptor, a large protein span-
ning the bilipid cellular membrane. Here, the receptor is
prepared with an electron on an donor site D (electronic
state D(cid:105), at an energy ED). This is illustrated in the
left-most portion of Figure 1. Detection involves ET to
an acceptor site A, at energy EA, which is ∆ below ED.
A direct D(cid:105) → A(cid:105) transition does not conserve energy,
and thus is unlikely. On the other hand, if an odorant
vibrational mode has quantization ωo = ∆, then an in-
elastic D(cid:105) → A(cid:105) ET may readily occur with the associ-
ated excitation of a quantum of vibration in the odorant.
A frequency-selective response of this type is referred to
as inelastic electron tunneling spectroscopy (IETS), and
an olfactory IETS process is depicted in the center part
of Figure 1. Then, the transferred electron is presumed
to trigger some subsequent process, which continues the
chain of events in sensing (in Figure 1, this ET triggers
the activation of the G protein, which releases an α sub-
unit).
In this paper, we focus on the D(cid:105) → A(cid:105) transition
without considering downstream and upstream events in
the sequence leading to an olfactory detection. A fully-
quantum, numerical model of the vibration-coupled IETS
mechanism is presented here.
9
1
0
2
r
a
M
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
7
2
1
0
1
.
1
1
8
1
:
v
i
X
r
a
2
FIG. 1. The inelastic electron tunneling spectroscopy (IETS) mechanism is illustrated in the context of human olfaction. A
human odorant receptor is known to be a G-protein coupled receptor (GPCR, shown in blue, with an associated G protein).
When an odorant molecule docks in the receptor (left figure), an electron transition from D (quantum state D(cid:105) at energy
ED) to A (state A(cid:105) at energy EA) requires a change in energy that is facilitated by the excitation of a quantum of molecular
vibrational energy. Once complete, the transferred electron may trigger the activation of the associated G protein, leading to
an action potential and a sensory detection event. Without an odorant having a properly-quantized vibrational spectrum, the
electron transition and subsequent detection are much less probable in the absence of other mechanisms by which the electron
can relax from ED to EA.
Indeed, previous models of the vibrational theory of
olfaction exist. Some of these models obtain an ET rate
through Fermi's golden rule and are semi-classical in the
sense that the electronic component is given a quantum
treatment, and the vibrational components are treated
classically.12,13 Other models go beyond Fermi's golden
rule and obtain an ET rate from a master equation that
includes the spectrum of the odorant and environmen-
tal harmonic oscillators.14 -- 17 In these models, the sys-
tem is partitioned such that the degrees of freedom of
the oscillators are traced over in the development of the
master equation. In one of these models, the dissipation
of energy (or power) was found to enhance a receptor's
selectivity for a particular vibrational quantization in an
odorant.15 Another such model of chiral odorants showed
that an energy difference between the ground states of
enantiomer pairs could also lead to a rate-based dissi-
pative discriminatory mechanism.18 While other mod-
els treat vibrational modes of the odorant and the en-
vironment as part of the reservoir and trace over these
degrees of freedom to obtain a master equation, we de-
velop a non-equilibrium model in which both the electron
transfer and the dominant odorant vibrational mode are
treated explicitly. This model lends itself to a calcula-
tion of power dissipation and highlights the link between
power dissipation and electron transfer.
II. MODEL
A. Framework
FIG. 2. The model developed in this paper includes the ex-
plicit, fully-quantum treatment of a D(cid:105) → A(cid:105) electron trans-
fer event (ET) coupled to a dominant odorant vibrational
mode. The "Electronic state" represents the ET event, and
the "Vibrational system" represents the dominant odorant
mode. This ET event is coupled to the thermal environment
both directly, and indirectly via the odorant vibrational mode.
In each case, the parameter that characterizes coupling be-
tween two systems is listed.
The framework for this fully-quantum model of the vi-
brational theory of olfaction is an electronic two-state
system dissipatively coupled in two ways to the ther-
mal environment: (1) a direct coupling to environmental
G proteinOlfactoryReceptorOdorant with excited vibrational modesDissociation of α subunit𝑒−𝑒−Odorant molecule(Acetophenone, C8H8O)Lipid bilayer𝐸𝐷𝐸𝐴Electronic stateVibrational systemEnvironmentT𝑇𝑒𝑇1λdegrees of freedom, and (2) and indirect coupling via
environmentally-damped odorant vibrations. The dy-
namics of the electronic system and one dominant odor-
ant vibrational mode are explicitly treated in this paper.
Dissipative effects are driven by interactions with the to-
tal environment, which is not explicitly modeled here but
may include receptor vibrations and solvent degrees of
freedom. The indirect dissipative path gives rise to spec-
troscopic behavior and is modeled using damped quan-
tum oscillators to treat odorant vibrational modes. The
direct electron-environment dissipation path is treated
by building electronic relaxation into the model. This is
depicted schematically in Figure 2. For computational
tractability, only one odorant vibrational mode is pre-
sumed to be strongly dominant (the coupling between
the electron and all other odorant modes is ignored).
The electronic state is described by the Hamiltonian
He = (∆/2)σz − γ σx, where σx and σz are Pauli oper-
ators, which we may write in terms of transition opera-
tors Pjk ≡ j(cid:105)(cid:104)k and projection operators Pj ≡ j(cid:105)(cid:104)j:
σx = PAD + PDA and σz = PD − PA. The parameter
γ is the hopping energy between the two states D(cid:105) and
A(cid:105), also known as the coupling constant, HAB, in quan-
tum chemistry. The detuning between the two states, ∆,
functions as the driving force behind the electron trans-
fer: ∆ = ED − EA.
a quantum harmonic oscillator, with Hamiltonian Hv:
The dominant odorant vibrational mode is modeled as
Hv =(cid:18)a
†
a +
1
2(cid:19) ωo =
P 2
2mo
+
1
2
moω2
o
Q2.
(1)
Here, ωo is the frequency of the dominant oscillator mode;
is the reduced Planck constant; and mo is the dominant
vibrational mode's effective mass. The position (coordi-
nate) operator Q and the momentum operator P may be
written in terms of a† and a, the creation and annihila-
tion operators, respectively:
Q =(cid:114)
P = i(cid:114) moωo
2
+ a(cid:1)
− a(cid:1) .
(2)
The right-hand side of Eqn. (1) is written in terms of
the kinetic energy ( P 2/2mo) and the potential energy
2moωo(cid:0)a
(cid:0)a
Q2/2(cid:17) operators for this vibrational mode.
The coupling between the odorant and the receptor's
electronic state is described by a linear coupling term in
the Hamiltonian,
(cid:16)moω2
and
†
†
o
Hev =
gev
2
σz Q.
(3)
The coupling constant gev depends on λ, the reorganiza-
tion energy of the odorant's dominant vibrational mode:
oλ. Thus, the fully-quantum system is de-
scribed by the total Hamiltonian, H, given by
gev =(cid:112)moω2
H = He + Hv + Hev
=
∆
2
σz − γ σx +
P 2
2mo
+
1
2
moω2
o
Q2 +
gev
2
σz Q.
(4)
3
The effects of the bath on the electron+odorant sys-
tem may be modeled by treating it as an open quantum
system in a Markovian environment. To do this, we use
the Lindblad equation,19
d
dt
ρ = −
i
(cid:104) H, ρ(cid:105) + D .
(5)
The density operator, ρ(t), describes the time-dependent
state of the electron+odorant system. The first term de-
scribes unitary dynamics and is equivalent to the quan-
tum Liouville equation. The dissipator, D, models envi-
ronmental effects and is given by
D =
†
Lj ρ L
j −
s(cid:88)j=1
†
j
1
2(cid:110) L
Lj, ρ(cid:111) .
(6)
The Lindblad operators, { Lj}, also known as environ-
mental channels, describe the effects of the environment
on the odorant-receptor complex.
In this model, the
indirect-path Lindblad operators are chosen so as to
damp only the vibrational subsystem. In particular, two
Lindblad operators will be used:
L1 =
1
√T1
a,
and
L2 = exp(cid:18)−
ωo
2kBT(cid:19) 1
√T1
†
a
.
(7)
The operator L1 removes energy from the vibrational
system, with T1 being a phenomenological characteristic
time for exponential energy relaxation via this indirect
path. After a time t (cid:29) T1, the system will relax to its
ground state if only L1 is used. The operator L2 excites
the system and includes a prefactor which depends on
temperature T . When both L1 and L2 are used in con-
cert, they drive the system to a Boltzmann distribution.
Finally, the direct electron-environment coupling is mod-
eled using an additional pair of Lindblad operators, L3
and L4, given by:
L3 =
PDA,
and
1
√Te
L4 = exp(cid:18)−
∆
2kBT(cid:19) 1
√Te
PAD.
(8)
Here, Te is the characteristic time for an exponential
relaxation via the direct electron-environment dissipation
path. The combination of L3 and L4 drives the system
to a Boltzmann distribution.
To quantify the strength of enviromental coupling, it
is helpful to define coupling strength ratios. For indirect
coupling, the harmonic oscillator's period To = 2π/ωo is
the natural time scale, so χv ≡ To/T1 provides a useful
measure of vibron-environment coupling. A vibrational
system decoupled from the environment is characterized
by χv → 0. For the electronic system, the detuning,
∆, is the natural energy scale, so we use χe ≡ t∆/Te to
characterize the strength of direct electron-environment
coupling, where t∆ ≡ /∆. Small χe characterizes an
electronic system for which direct environmental dissi-
pation is suppressed. Additionally, when comparing the
two coupling strengths, we use the ratio ξ ≡ Te/T1, for
which ξ → 0 when direct electron-environment coupling
is dominant, and ξ → ∞ for dominant indirect electron-
environment coupling via the odorant vibrational mode.
In order to qualitatively show that in the regime of in-
terest, the fully-quantum model is preferable over a semi-
classical treatment, we reduce the fully-quantum model
to a semi-classical treatment. If the kinetic energy of the
oscillator is ignored, and if the coordinate is treated as a
classical variable instead of as an operator ( Q → Q), then
the Hamiltonian reduces to a Marcus-type Hamiltonian,
H (M ), which acts on the Hilbert space of the electronic
system only:
H (M ) =
∆
2
σz − γ σx +
1
2
gev σxQ2 +
1
2
moω2
oQ2.
(9)
B. Choice of Parameters
In the absence of concrete experimental data for model
parameters, we pick typical values for biological sys-
tems and theorized values from the literature. Thus,
all calculations are performed at ambient temperature
(T = 293 K) unless the role of temperature is inves-
tigated. A tunneling energy γ = 1 meV is chosen,
in following with analysis found in the literature.12 D
and A are treated as single molecular orbitals coupled
to each other by a weak hopping integral γ, as there
should be essentially no tunneling from D to A in the
absence of the odorant or any other dissipative pathway.
We often use a D-A detuning of ∆ ∼ 200 meV (1613.6
cm−1), chosen since the interesting range in olfaction is
70 meV < wo <400 meV ( 564.77 cm−1 < wo < 3227.3
cm−1).12,20 The odorant reogranization energy is varied
between 30 meV (∼ kBT at biological temperatures) and
300 meV (∼ ∆).
Parameter
Value
∆
γ
λ
T
200 meV ( 1613.6 cm−1)
1 meV
30-300 meV
293 K
TABLE I. Physically-interesting values of model parameters
are chosen in following with treatments from the literature
and are enumerated here.12,20
4
(a) λ = 30 meV
(b) λ = 300 meV
FIG. 3. A fully-quantum treatment within the vibrational
theory of olfaction is justified. The eigenvalues of the fully-
quantum Hamiltonian H are plotted relative to the Marcus-
type potential energy surfaces V (M )
(Q) vs. Q. This plot is
shown for two different values of the odorant reorganization
energy λ: once with λ = 30 meV and again with λ = 300 meV.
In each case, the energy discretization of fully-quantum sys-
tem is significant compared to feature sizes of semi-classical
adiabatic Marcus-type potential surfaces. This suggests that
treatment using the fully-quantum model will lead to behav-
iors not captured by a Marcus-type semi-classical model.
k
C.
Justification for a Fully-quantum Treatment
by solving the time-independent Schrodinger equation:
A comparison of the statics between the fully-quantum
treatment and the semiclassical Marcus-type reduction
reveals the necessity for a fully-quantum treatment of
IETS within this model. The stationary states φk(cid:105) and
eigenvalues Ek for the Hamiltonian of Eqn. (4) are found
H φk(cid:105) = Ek φk(cid:105)
(10)
The semiclassical H (M ) has two eigenstates, {φ(M )
with k ∈ 1, 2, and two eigenvalues V (M )
(cid:105)},
, which solve the
k
k
-4-3-2-101234-40-20020406080100120140-4-3-2-101234-100-80-60-40-20020406080time-independent Schrodinger equation:
k
k
k
H (M )(cid:12)(cid:12)(cid:12)φ(M )
k (cid:69) = V (M )
(Q) and eigenstates φ(M )
(cid:12)(cid:12)(cid:12)φ(M )
k (cid:69)
Since H (M ) is a function of the coordinate Q, so also
are the eigenvalues V (M )
(Q)(cid:105).
k
The eigenvalues V (M )
(Q) provide an adiabatic potential-
energy landscape for the semi-classical system. Figure
3 shows the potential energy surfaces defined by the
{V (M )
(Q)} for the semiclassical system, along with the
lowest-energy eigenvalues for the fully-quantum system.
The {V (M )
(Q)} surfaces are plotted with Q scaled to
Q0 = gev/2moω2
o. The energy quantization in the fully-
quantum treatment is significant compared to the feature
sizes seen in the potential landscape of the semi-classical
treatment. This indicates that the fully-quantum treat-
ment of the electron+odorant system is justified,
in-
deed. Thus, the fully-quantum treatment will yield re-
sults which semi-classical models cannot capture. This
comparison is performed for two different values of λ. It
is seen that for a larger λ, the potential barrier between
the coordinates Q = ±1 becomes more significant.
k
k
prepared in an initial state ρ(t ≤ 0), which is chosen to
be, ρth
0 , the thermal equilibrium density matrix for the
electron+ odorant system:
(11)
5
with
ρ(0) = ρth
0 ≡
1
Z
exp(cid:18)−
1
kBT
H0(cid:19) ,
Z = Tr (cid:18)exp(cid:18)−
1
kBT
H0(cid:19)(cid:19) .
D
Here, H0 is the similar to H from Equation (4), but dif-
fers in that we set ∆(t ≤ 0) = ∆0. ∆0 is a large negative
potential applied for the sole purpose of confining the
electronic state to D(cid:105). Then, at t = 0, ∆(t) is changed
abruptly to a positive, static value appropriate to the
physics of the receptor. Next, ρ(t) is obtained for t > 0
by solving Eqn. (5) numerically using the new, constant
H for t > 0, and using ρ(0) as the initial value. The
probabilities for finding the electron on the donor site
and the acceptor sites, PD and PA, respectively, are cal-
culated as the expectation value of the projection opera-
tors: PD = (cid:104) PD(cid:105) and PA = (cid:104) PA(cid:105).
Figure 4 shows non-equilibrium model data from which
an electron transfer time tET is calculated. tET is defined
as the time when PD drops from PD(0) (cid:39) 1 to a threshold
value P (thresh)
. The electron transfer rate k is the recip-
rocal of the electron transfer time: k = 1/tET . While
any P (thresh)
< 0.5 could serve as a threshold, we choose
P (thresh)
= 0.2 because this is low enough to preclude
D
complicating oscillations in PD(t) as PD(t) → P (thresh)
,
and yet high enough to allow reasonable calculation
times. This calculation is performed with large ξ so that
indirect dissipation is dominant. Enhancing the direct
dissipation pathway (reducing ξ) would allow a faster re-
laxation, resulting in a faster ET and a higher rate k.
The Rabi oscillation frequency γ/π sets the upper
speed limit for quantum charge transport in this system
in the absence of coherent driving, so half of the Rabi
oscillation period sets the lower limit for physical electron
transfer times: tET > π/2γ. Here, with P (thresh)
= 0.2,
tET = 37.5 ps, which satisfies tET > π/2γ (cid:39) 1 ps. The
result for tET is plausible because tET does not violate
its lower limit π/2γ and the result is well below the
biological time scale of ms for actuating GPCRs.21
D
D
D
E. Calculation of power dissipation, ¯pdiss
Power dissipation may be calculated within this model.
Power dissipation here is given by
pdiss = −pE,ev − pE,v − pE,e ,
(12)
D. Calculation of electron transfer time, tET
FIG. 4. Calculation of threshold defined electron transition
time, tET , from the dynamic solution to Eqn. (5). Here, ρ(t)
is calculated and the probabilities PD and PA for finding the
electron on D or A are calculated.
tET is defined as the
time when PD drops from PD(0) (cid:39) 1 to a threshold value
P (thresh)
= 0.2, the result is tET = 37.5 ps.
Dashed black lines indicate the threshold P (thresh)
= 0.2 and
the electron-transfer time, t = tET , defined as the time that
PD drops below P (thresh)
. When P (thresh)
D
D
D
.
D
Electron transfer times, tET , within the fully-quantum
treatment are obtained from the time dynamics of the
non-equilibrium model calculations. First, the system is
00.010.020.030.040.050.0600.20.40.60.81where
6
pE,ev = Tr (cid:16)D Hev(cid:17) ,
pE,v = Tr (cid:16)D Hv(cid:17) , and
pE,e = Tr (cid:16)D He(cid:17) .
(13)
We interpret pE,ev + pE,v + pE,e as the rate of work done
on the electron+odorant system by the environment.
Average power dissipation, ¯pdiss, is evaluated over the
electron transfer time tET :
1
tET (cid:90) tET
0
¯pdiss =
III. RESULTS
pdiss(s) ds .
(14)
Results are presented in three different regimes: (A)
in the limit of dominant indirect coupling (weak direct
electron-environment coupling, ξ → ∞); (B) in the limit
of dominant direct electron-environemtn coupling (weak
indirect electron-phonon-environment coupling, ξ → 0);
and (C) an intermediate regime, in which both couplings
are present, but neither is dominant.
A.
Indirect Coupling Dominates
When indirect coupling provides the dominant dissipa-
tion pathway, a rich set of spectroscopic behaviors is seen.
Here, we explore the relationship between rate and odor-
ant frequency, odorant reorganization energy, tunneling
energy and temperature, and power dissipation.
1. Resonant Peaks in ET Rate
When indirect coupling is dominant (ξ → ∞), electron
transfer rates exhibit resonant peaks. Figure 5 shows the
ET rate k as a function of odorant vibrational angular
frequency ωo = 2πfo. The frequency axes is scaled in
units of ∆ = ED − EA. Resonant peaks in k(fo) occur
at frequencies ωo = ∆/s, where s is a positive integer.
As odorant environmental coupling (χv) increases, the
spectral peaks broaden.
The resonance of Figure 5 can be understood in the
limit of λ = 0, γ = 0 and T = 0. Here, the eigen-
states of the system are product states X(cid:105)⊗n(cid:105) = Xn(cid:105),
with eigenenergies EX + ωo (n + 1) /2 for occupation
number n ∈ {0, 1, 2, . . .} and X ∈ {D, A}. This spec-
trum is depicted in Figure 6. The initial state is ap-
proximately D0(cid:105), with energy ED + ωo/2.
If ∆ =
ED − EA = sωo for some positive integer s, then the
initial energy ED + ωo/2 matches exactly the eigenen-
ergy EA + ωo (s + 1) /2, facilitating an D → A electron
transfer event from the combined state D0(cid:105) to As(cid:105). In
FIG. 5. The electron transfer rate, k, from state D(cid:105) to state
A(cid:105) exhibits resonant tunneling and environmentally-driven
broadening of peaks. Here, peaks occur under a resonant
condition that arises when the detuning ∆ = ED − EA is
equal to an integer multiple of the vibrational quantization
ωo. Additionally, resonant peaks are broadened as odorant-
environment coupling increases, as quantified by χv.
order for the system to settle to A0(cid:105) and for the elec-
tron transfer to "complete" (PD → 0), the energy sωo
must be dissipated. Thus, low s minimizes the vibra-
tional energy dissipation and time required for the ET
event, and peaks in the rate k grow larger with decreasing
s, with s = 1 providing the highest peak. Multi-phonon
processes -- infrequently-considered in the theory of olfac-
tory IETS -- could play a role. In this model, ET rates
are increased when the energy of fewer phonons must be
dissipated to allow the system to settle to A0(cid:105).
2. Odorant Reorganization energy, λ
The absence of the odorant is modeled in the limit of
λ → 0. Since coupling between the ET event and the
environment already is negligible, the electronic system
is unable to dissipate energy. Thus, the D(cid:105) → A(cid:105) tran-
sition becomes less likely because of the non-negligible
detuning ∆, and rate decreases to zero. This is seen in
Figure 7. On the other hand, increasing λ strengthens the
electron-vibration coupling. An excessively strong cou-
pling (high λ) broadens the peaks in k(ωo/∆). Figure
7 shows that increasing the reorganization energy λ con-
tributes to a higher electron transfer rate k for lower fre-
quencies. Resonant peaks are broadened as λ increases.
The λ-dependence of rate k seen here is appears to be
consistent with Fermi's golden rule.
00.511.510810910107
(a) χv = 1/2
FIG. 6. The eigenenergy spectrum for the system results
in a resonant effect for the ET rate k.
In the limit of
γ = λ = T = 0, the eigenstates are Xn(cid:105), with energies
EX + ω (n + 1/2). When ∆ = ED − EA = sω for some
integer s, a resonant transition D0(cid:105) → As(cid:105) is favorable, as
is the resonant back-transitionAs(cid:105) → D0(cid:105). The probability
PD does not decrease to a small value until the system can
dissipate energy for the state to approach A0(cid:105). Since this
dissipation takes time, the minimum non-zero value (n = 1)
enables the fastest dissipation and thus the fastest ET time
and the highest ET rate k.
3. Tunneling energy, γ
The results of Figure 8 shows that increasing the tun-
neling energy γ enables a higher electron transfer rate.
This is consistent with the fact that the upper speed
limit of electron transfer is the Rabi oscillation frequency
πγ/. Additionally, the γ-dependence of rate k seen here
appears to be consistent with Fermi's golden rule.
4. Temperature, T
(b) χv = 2
Figure 9 shows ET rate k as a function of temper-
ature.
In the zero-temperature limit, a non-zero k re-
veals the quantum nature of the system, as a classical or
semi-classical prediction for the ET rate would be zero
at in the absence of thermal excitation. For terrestrial
and biological temperatures T < 300 K, ET rate k is
largely constant. As T increases to high temperatures
(kBT ∼ ωo), k begins to decrease. This is because as
T increases, the system is increasingly thermally excited,
raising the probability PD(t) of finding the electron on
the D site. This increases the time tET it takes for PD(t)
to decay to the threshold defined in our calculation, thus
decreasing k. While the temperature-dependence of k
from this model is largely consistent with Fermi's golden
rule, a point of divergence between the two models is
rate behavior at low temperature. Fermi's golden rule
predicts k = 0 at T = 0; however, the present model
predicts k (cid:54)= 0 at T = 0.
Interestingly,
a
quantum-to-classical
transition
FIG. 7. Resonant peaks in k(fo) are broadened as odorant
reorganization energy, λ, is increased. Also, increasing λ con-
tributes to a higher ET rate, k, for lower frequencies. This is
seen in subfigure (a) for a weaker odorant-environment cou-
pling [χv = 1/2], and again in subfigure (b) for a stronger
coupling [χv = 2].
emerges when we consider how temperature T and
oscillator frequency ωo affect rate k. Figure 10 shows
calculations of k for various values of ωo and T . For low
ωo, reducing T decreases the ET rate k. This result is
consistent with classical and semi-classical rate models,
which require thermal excitation to cross a potential
barrier. For low ωo, the fully-quantum system gives
rise to this semi-classical behavior because lowering
ωo reduces the spacing between eigenvalues of the
fully-quantum system (see Figure 3). For low-enough
ωo, the energy quantization in the fully-quantum model
is small compared to the barrier height between states
EnergyED+¯hωo(cid:0)12(cid:1)D0iED+¯hωo(cid:0)1+12(cid:1)D1iED+¯hωo(cid:0)2+12(cid:1)D2i......AsiEA+¯hωo(cid:0)s+12(cid:1)...A1iEA+¯hωo(cid:0)1+12(cid:1)A0iEA+¯hωo(cid:0)12(cid:1)s¯hωo00.511.52108109101000.511.5210910108
(a) χv = 1/2
FIG. 9. Electron transfer rate, k, from state D(cid:105) to state
A(cid:105) is largely temperature-independent at terrestrial temper-
atures and below. ET rate is shown here as a function of
environmental temperature T for different values of odorant-
environment coupling χv. When T is sufficiently large that
kBT ∼ ω, ET rate decreases since thermal environmental
fluctuations enable A(cid:105) → D(cid:105) excitation.
sponse to isotopomers. Specifically, we considered a re-
ceptor's ability to spectroscopically distinguish acetophe-
none from its fully-deuterated isotopomer acetophenone-
d8. Here, we assume that the C-H and C-D stretch play
the distinguishing role.2,9
o
o
o
Figure 11 shows the spectroscopic response of k for
two receptors: one receptor is resonantly tuned to the
C-H stretch frequency f (C−H)
= 2300 cm−1 by setting
∆ = ∆C−H = ω(C−H)
, and another is resonantly tuned
to the C-D stretch frequency f (C−D)
= 3000 cm−1 by
o
setting ∆ = ∆C−D = ω(C−D)
. The response of each
receptor is calculated for various values of ωo, and the
strongest resonant peak occurs as expected for the pre-
ferred odorant when ∆ = ωo. For each receptor, if
odorant-environment coupling is weak (χv = 1/2), ET
rate k drops by almost two orders of magnitude when
the non-preferred isotopomer is in the receptor. For a
stronger odorant-environment coupling (χv = 2), spec-
tral broadening results in the spectroscopic k(ωo) re-
sponse, and rate drops by more than 50% when the pre-
ferred molecule is replaced by its isotopomer. Increased
environmental coupling reduces the receptor's selectivity
to its preferred odorant.
(b) χv = 2
FIG. 8. Increase in tunneling energy, γ, contributes to higher
electron transfer rate, k, from state D(cid:105) to state A(cid:105). This is
shown for weak odorant-environment coupling, χv = 1/2 and
for strong odorant-environment coupling, χv = 2.
D(cid:105) and A(cid:105) in the semi-classical system. Thus, the
semi-classical behaviors of the system will be dominant.
On the other hand, when ωo is large, energy quantization
in the system is significant, the system manifests more
quantum mechanical behaviors. Here, increasing T only
slightly decreases k, as discussed previously. With large
ωo, a high rate k persists, even at low temperature,
T = 0. This is a truly non-classical result.
5. Discrimination between isotopomers
Next, the model developed here is applied to the study
of receptor selectivity in the odorant-receptor complex re-
The dissipation of power and energy is essential to a
D(cid:105) → A(cid:105) transition. A plot of average power dissipation
¯pdiss (see Figure 12) for various odorant frequencies shows
that the system dissipates the highest rates of average
6. Power dissipation
00.511.51081010101200.511.521071081091010101110120500100015002000250000.511.522.5310109
(a) χv = 1/2
(b) χv = 2
FIG. 11. The model developed here exhibits a differential
response to isotopomers.
(a) A receptor tuned to the C-H
stretch responds to the C-H stretch with a rate k that is more
than one order of magnitude higher than the rate due to the
presence of a C-D stretch vibrational mode. A receptor tuned
to the C-D stretch exhibits a similarly higher ET rate when
an odorant with a C-D ligand is present than when an odorant
with a C-H ligand is present. (b) Peak broadening occurs with
stronger odorant-environmental coupling, and the change in
rate between isotopomers is reduced for each receptor.
ence of a C-H (or C-D) bond and divide by the power
dissipation in the presence of a C-D (or C-H) stretch-
ing mode. With an odorant-environment coupling of
χv = 1/2, selectivity for the preferred odorant is ∼ 15 dB
(see Table II). Stronger odorant-environment coupling
(χv = 2) broadens the peaks in the ¯pdiss(fo) curve, and
selectivity drops to ∼ 5 dB (see Table III).
FIG. 10. Model ET rates exhibit both quantum and
classical/semi-classical behaviors over the spectrum of odor-
ant frequencies fo. In the low frequency limit (ωo (cid:28) ∆), in-
creasing temperature increases the rate k in a behavior consis-
tent with a classical or semi-classical system in which thermal
excitation allows the system to surmount a classical barrier.
In the high-frequency limit (ωo (cid:29) ∆), ET rate, k, decreases
with a significant increase in temperature. This is consistent
with the behavior noted in Section III A 4.
power at frequencies satisfying ∆ = sωo for some pos-
itive integer s. This coincides exactly with the resonant
peaks in rate (see Figure 5), underscoring the enabling
role dissipation plays in this process: the more quickly
the electron can dissipate power, the faster it can make
the D(cid:105) → A(cid:105) transition, increasing the ET rate. This
relationship is a direct consequence of the conservation
of energy, and the ET is thus facilitated by its ability to
relax by ∆ via the excitation of odorant phonons or by
dissipating power to the thermal environment.
The relationship between rate k and ¯pdiss exhibits a
highly linear relationship, as seen in the scatter plot of
Figure 13. This scatter plot shows the correlation be-
tween the rate data from Figure 5 and the power dissipa-
tion data of Figure 12. One regression line is remarkably
effective for three different odorant-environment coupling
strengths (χv ∈ {0.5, 1, 2}).
Finally, we use power dissipation to quantify the se-
lectivity of a receptor within this model to its preferred
isotopomer. Power dissipation ¯pdiss is plotted as a func-
tion of odorant frequency fo in Figure 14. The power
dissipation curves of Figure 14 are very similar to the
rate-versus-frequency curves for the same receptors (see
Figure 11).
We quantify receptor selectivity by taking the ratio
of power dissipation at the characteristic vibrational fre-
quency of the preferred odorant to the rate of power
dissipation at the frequency characteristic of the non-
preferred odorant. For the receptor tuned to the C-H (or
C-D) stretch mode, we use power dissipation in the pres-
00.511.5210810910100100020003000400050001061081010010002000300040005000107108109101010
FIG. 12. Power dissipation ¯pdiss to the environment exhibits
a frequency-dependence very similar to the the frequency de-
pendence of rate k to odorant frequency (compare with Fig-
ure 5). Peaks for ¯pdiss occur when ∆ = sωo for some in-
teger s, and resonant peaks broaden with stronger odorant-
environment interaction (increasing χv).
(a)
χv = 1/2
FIG. 13. Power dissipation ¯pdiss exhibits a linear correlation
with ET rate. This scatter plot of rate data from Figure 5
and power dissipation data from Figure 12 shows a highly
linear relationship between the two sets of data. A fitting line
is drawn in purple, and the data lies along this same line for
several odorant-environment coupling strengths.
B. Direct (Electron-environment) Coupling Dominates
When direct electron-environment coupling provides
the strongly-dominant dissipation path (ξ → 0), spectro-
scopic behaviors vanish, and rate becomes largely inde-
pendent of the vibrational frequency of the odorant. This
is seen in Figure 15. Here, a raised electron-environment
(b)
χv = 2
FIG. 14. Average power dissipation to the environment
is frequency-selective -- and therefore ligand-selective. Power
dissipation peaks when a receptor tuned to the C-H (or C-D)
stretch couples to a ligand with a frequency fo matching the
frequency of the C-H (or C-D) stretch.
coupling (increasing χe) enables a faster relaxation and
a higher ET rate k.
A very subtle resonant effect in rate is seen if χe is low.
Slight deviations in the ET rate at resonant frequencies
(ωo = s∆/ for some integer s > 0) are revealed (see
data for χe = 0.01). These deviations are due to a tran-
sient response in which a resonant condition arises, and
the conservation of energy facilitates rapid power transfer
from electron to an integer number of odorant phonons.
This allows a faster ET: the probability PD(t) crosses the
threshold P (thresh)
sooner, resulting in a slightly higher
rate k. However, since this energy cannot be dissipated
D
00.511.52100101102103010020030040050060000.511.522.53101001000200030004000500010010201000200030004000500010-1100101102Receptor
tuned to
Power dissipa-
tion (pW)
C-H
Stretch
C-D
Stretch
Selectivity
(dB)
C-H stretch 521.7
C-D stretch 10.6
17.2
420.7
14.81
16
TABLE II. Power dissipation from the receptor-odorant com-
plex to the environment is enhanced when a ligand is present
for which the receptor is tuned (i.e., ∆ = ωo). Here, an
odorant-environment coupling of χv = 1/2 is assumed. When
the preferred ligand is replaced by its related isotopomer,
∆ (cid:54)= ωo, and power dissipation is suppressed by ∼ 15.4 dB.
This frequency-selective dissipation is used to quantify the se-
lectivity of the receptor to a particular vibrational mode over
in a related isotopomer.
Receptor
tuned to
Power dissipa-
tion (pW)
C-H
Stretch
C-D
Stretch
Selectivity
(dB)
C-H stretch 133.4
C-D stretch 31.4
46.1
106.9
4.62
5.32
TABLE III. In analogy to Table II, receptor selectivity is cal-
culated and tabulated for stronger odorant-environment cou-
pling (χv = 2). Increased odorant-environment coupling re-
duces receptor selectivity (compare to Table II).
D
Figure
16
from the odorant to the environment, the energy is sub-
sequently transferred back to the electron, with some
A(cid:105) → D(cid:105) back-transfer, and a PD(t) that had crossed
below P (thresh)
electron-to-
environment interaction indeed drives an exponential
relaxation characterized by time constant Te. Here,
may cross above P (thresh)
the direct
shows
plotted with an exponential fit, given by
E(t) = Tr (cid:16) H ρ(t)(cid:17), the expectation value of energy, is
fE(t) = E(∞) + (E(0) − E(∞))(cid:16)1 − e
−t/Te(cid:17) .
The calculated E(t) from the model matches the expo-
nential fE(t) very precisely.
D
again.
that
(15)
C. Both Direct and Indirect Coupling are Significant
Here, we explore the case in which neither the di-
rect nor indirect dissipative pathways are negligible.
In Figure 17, the resonant peaks in the rate k be-
come less pronounced as the system transitions from
an indirect-coupling-dominated (larger ξ) to direct-
coupling-dominated (smaller ξ). This is because the ad-
ditional (direct) dissipative pathway assists the electron
transfer, raising the ET rate most noticeably in the non-
resonant regions (∆ (cid:54)= sω0 for positive integers s) where
ET was most highly suppressed with high ξ (compare
11
FIG. 15. Resonant behaviors are almost completely sup-
pressed when the indirect dissipation path is disrupted by
severing the odorant-environment connection (χv → 0). Here,
the rate is largely independent of odorant vibrational fre-
quency f0 and is determined solely by the strength of di-
rect dissipation χe. ET rate increases with increasing direct
electron-environment coupling χe.
FIG. 16. The direct electron-environment dissipation path
drives an exponential relaxation with time constant Te. Here,
Te = 33 fs, and electron+odorant total energy E(t) matches
an exponential fit fE(t).
Figure 5 to the red and blue traces of Figure 17). If ξ be-
comes small enough, the spectroscopic response of k(f0)
vanishes as the direct dissipative pathway dominates and
the resonant behavior of the indirect dissipation path is
masked.
The degradation of the receptor spectroscopy is also
seen in the dissipation plots of Figure 18. For high ξ, the
spectroscopic behaviors are readily seen in strong peaks
in the dissipation curve ¯pdiss(fo). As ξ is reduced, spectro-
00.511.52101210131014024681040608010012014016018012
(a) Receptor tuned to C-D
(b) Receptor tuned to C-H
FIG. 18.
Spectroscopic effects in power dissipation are
masked with increasing direct-path dissipation (decreasing ξ)
for receptors tuned to a specific isotopomer. Subfigure (a) is
for a receptor tuned to the C-D stretch frequency, and (b) is
for a receptor tuned to the C-H stretch frequency.
Receptor tuned to C-H stretch
Indirect to direct Selectivity
coupling ratio, ξ
(dB)
6 × 107
1 × 103
5 × 102
14.8
3.0
1.5
FIG. 17. As the strength of direct coupling is increased (ξ
decreases), the resonant peaks broaden, and spectroscopic be-
haviors in the ET rate become masked.
scopic behaviors are less pronounced as direct-path dissi-
pation begins to compete with the indirect-path dissipa-
tion, enabling higher levels of dissipation in previously-
suppressed (non-resonant) frequencies. Eventually, the
spectroscopic behavior is completely masked as ξ is fur-
ther reduced.
The loss of spectroscopic behaviors with reduced ξ can
be quantified in the selectivities of the receptors tuned to
either the C-D or C-H stretch modes. Selectivity for these
receptors is listed in Tables IV and V for the case where
χv = 1/2. High values of ξ yield the receptor selectivities
seen in the indirect-path-dominant regime (see Table II).
When ξ is reduced enough, selectivity is degraded so that
isotopomers are no longer distinguishable (selectivity is
below ∼ 3 dB).
Receptor tuned to C-D stretch
Indirect to direct Selectivity
coupling ratio, ξ
(dB)
6 × 107
2 × 103
1 × 103
16.0
3.0
1.5
TABLE IV. Strengthening the direct electron-environment
dissipation path (reducing ξ) degrades the sensitivity of a re-
ceptor tuned to the C-D stretch frequency.
IV. CONCLUSIONS
A numerical model was presented for olfactory IETS
in which the two-state electronic system was coupled to
a thermal environment in two ways: (1) directly, and
TABLE V. Strengthening the direct electron-environment dis-
sipation path (decreasing ξ) degrades the sensitivity of a re-
ceptor tuned to the C-H stretch frequency.
(2) indirectly via a quantum harmonic oscillator mod-
00.511.521081091010101101000200030004000500010-2100102010002000300040005000100102eling the environmentally-damped dominant vibrational
mode of the odorant molecule. This model treats ex-
plicitly a dominant odorant vibrational mode and pro-
vides an method for calculating dissipation from the elec-
tron to the environment via both pathways. Resonances
were observed in the ET rate for odorant frequencies
ωo = ∆/s when indirect electron-environment coupling
is dominant, consistent with the vibrational theory of ol-
faction. The resonance in the ET process is related to the
system's ability to dissipate energy to the environment
via the indirect pathway only. In this limit, the model
also demonstrates a transition between semi-classical and
quantum behavior based on the size of the quantization
ωo. Odorants with small vibrational quantization ωo
exhibit an ET rate behavior in which decreasing temper-
ature reduces the ET rate, consistent with classical and
semi-classical ET rate models where thermal excitation
provides the means to pass over a potential energy bar-
rier. On the other hand, when ωo ∼ ∆, odorant quanti-
zation becomes significant, and reducing the temperature
of the reservoir has less effect on ET rate because quan-
tum mechanical tunneling becomes a more dominant ef-
fect. Also, this model predicts significant ET rates even
at T = 0 K, contrary to Fermi's golden rule. Behav-
ior like this is unique to quantum mechanical models.
Spectroscopic behaviors in the ET rate become masked
as direct electron-enviornment coupling plays a stronger
role in the ET. Nonetheless, the spectroscopic behaviors
may be relevant in some biological systems for which the
direct dissipative path is negligible.
The vibrational theory of olfaction is an interesting ap-
plication of quantum mechanics that requires further ex-
perimental techniques and developments for validation.
Experimental work will be required to validate this and
other models of IETS, and even the vibrational theory of
olfaction itself. No experiment has conclusively demon-
strated a spectroscopic olfactory mechanism, nor identi-
fied the existence of the D and A sites, nor eliminated
perireceptor events in distinguishing isotopomers. Fur-
ther developments in the model could incorporate the in-
clusion of additional odorant vibrational modes, as well
as a treatment of chirality in enantiomers. Models like
this may also be used in exploring spectroscopic effects
in GCPR more broadly in mammalian nervous systems
beyond olfactory receptors.22
ACKNOWLEDGMENT
The authors thank Craig S. Lent of the University of
Notre Dame for discussion on the topics of open quantum
systems applied to this model. This work was supported
by Baylor University under a new-faculty-startup grant.
1A. Keller and L. Vosshall, "A psychophysical test of the vibration
theory of olfaction," Nature Neurosci., vol. 7, no. 4, pp. 337 -- 338,
April 2004.
2M. Franco, L. Turin, A. Mershin, and E. Skoulakis, "Molecular
vibration-sensing component in drosophila melanogaster olfac-
13
tion," P. Natl. Acad. Sci. USA, vol. 108, no. 9, pp. 3797 -- 3802,
March 2011.
3S. Gane, D. Georganakis, K. Maniati, M. Vamvakias, N. Ragous-
sis, E. Skoulakis, and L. Turin, "Molecular vibration-sensing
component in human olfaction," PLoS One, vol. 8, no. 1, p.
e55780, 2013.
4E. Block, S. Jang, H. Matsunami, S. Sekharan, B. Dethier,
M. Ertem, S. Gundala, Y. Pan, S. Li, Z. Li, S. Lodge, M. Ozbil,
H. Jiang, S. Penalba, V. Batista, and H. Zhuang, "Implausibility
of the vibrational theory of olfaction," P Natl Acad Sci USA, vol.
112, no. 21, pp. E2766 -- E2774, 2015.
5L. Turin, S. Gane, D. Georganakis, K. Maniati, and E. Skoulakis,
"Plausibility of the vibrational theory of olfaction," P Natl Acad
Sci USA, vol. 112, no. 25, p. E3154, 2015.
6E. Block, S. Jang, H. Matsunami, V. S. Batista, and H. Zhuang,
"Reply to turin et al.: vibrational theory of olfaction is implau-
sible," Proc. Natl. Acad. Sci. US, vol. 112, no. 25, pp. E3155 --
E3155., 2015.
7M. Paoli, A. Anesi, R. Antolini, G. Guella, G. Vallortigara, and
A. Haase, "Differential odour coupling of isoptomers in the hon-
eybee brain," SCI REP-UK, vol. 6, p. 21893, 2016.
8E. Drimyli, A. Gaitanidis, K. Maniati, L. Turin, and E. M. Sk-
oulakis, "Differential electrophysiological responses to odorant
isotopologues in drosophilid antennae," eNeuro, pp. ENEURO --
0152, 2016.
9J. Brookes, A. Horsfield, and A. Stoneham, "The swipe card
model of odorant recognition," Sensors, vol. 12, pp. 15 709 --
15 749, 2012.
10L. Turin, "A spectroscopic mechanism for primary olfactory re-
ception," Chem Senses, vol. 21, pp. 773 -- 791, 1996.
11R. Hoehn, D. Nichols, H. . Neven, and S. Kais, "Status of the
vibrational theory of olfaction," Frontiers Phys, vol. 6, p. 25,
2018.
12J. Brookes, F. Hartoutsiou, A. Horsfield, and A. Stoneham,
"Could humans recognize odor by phonon assisted tunneling?"
Phys Rev Lett, vol. 98, p. 038101, January 2007.
13I. Solv'yov, P.-Y. Chang, and K. Schulten, "Vibrationally as-
sisted electron transfer mechanism of olfaction: myth or reality?"
Phys. Chem. Chem. Phys., vol. 14, pp. 13 861 -- 13 871, 2012.
14E. Bittner, A. Madalan, A. Czader, and G. Roman, "Quantum
origins of molecular recognition and olfaction in drosophila," J.
Chem. Phys., vol. 137, p. 22A551, 2012.
15A. Checinska, F. Pollock, L. Heaney, and A. Nazir, "Dissipation
enhanced vibrational sensing in an olfactory molecular switch,"
J. Chem. Phys., vol. 142, p. 025102, 2015.
16A. Tirandaz, F.T. Ghahramani, and V. Salari, "Validity exami-
nation of the dissipative quantum model of olfaction," SCI REP-
UK, vol. 7, p. 4432, 2017.
17J. Brookes, "Quantum effects in biology: golden rule in enzymes,
olfaction, photosynthesis and magnetodetection," Proc. R. Soc.
A, vol. 473, no. 20160822, 2017.
18A. Tirandaz, F. Ghahramani, and A. Shafiee, "Dissipative vibra-
tional model for chiral recognition in olfaction," Phys. Rev. E.,
vol. 92, p. 032724, 2015.
19G. Lindblad, "On the generators of quantum dynamical semi-
groups," Commun. Math. Phys., vol. 48, pp. 119 -- 130, 1967.
20I. Khemis, N. Mechi, and A. Lamine, "Stereochemical study of
mouse muscone receptor mor215-1 and vibrational theory based
on statistical physics formalism," Prog. Biophys. Mol. Bio., vol.
136, pp. 54 -- 60, 2018.
21J.-P. Vilardaga, M. Bunemann, C. Krasel, M. Castro, and
M. Lohse, "Measurement of the millisecond activation switch of
g protein -- coupled receptors in living cells," Nature Biotechnol,
vol. 21, pp. 807 -- 812, June 2003.
22R. Hoehn, D. Nichols, H. Neven, and S. Kais, "Neuroreceptor
activation by vibration-assisted tunneling," Sci Rep, vol. 5, p.
9990, 2014.
|
1310.8267 | 1 | 1310 | 2013-10-30T18:56:21 | Realization of Morphing Logic Gates in a Repressilator with Quorum Sensing Feedback | [
"physics.bio-ph",
"cs.ET",
"nlin.AO",
"q-bio.MN"
] | We demonstrate how a genetic ring oscillator network with quorum sensing feedback can operate as a robust logic gate. Specifically we show how a range of logic functions, namely AND/NAND, OR/NOR and XOR/XNOR, can be realized by the system, thus yielding a versatile unit that can morph between different logic operations. We further demonstrate the capacity of this system to yield complementary logic operations in parallel. Our results then indicate the computing potential of this biological system, and may lead to bio-inspired computing devices. | physics.bio-ph | physics | Realization of Morphing Logic Gates in a Repressilator with
Quorum Sensing Feedback
Vidit Agarwal, Shivpal Singh Kang and Sudeshna Sinha
Indian Institute of Science Education and Research (IISER) Mohali,
Knowledge City, SAS Nagar, Sector 81,
Manauli PO 140 306, Punjab, India
Abstract
We demonstrate how a genetic ring oscillator network with quorum sensing feedback can operate
as a robust logic gate. Specifically we show how a range of logic functions, namely AND/NAND,
OR/NOR and XOR/XNOR, can be realized by the system, thus yielding a versatile unit that can
morph between different logic operations. We further demonstrate the capacity of this system to
yield complementary logic operations in parallel. Our results then indicate the computing potential
of this biological system, and may lead to bio-inspired computing devices.
3
1
0
2
t
c
O
0
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
7
6
2
8
.
0
1
3
1
:
v
i
X
r
a
1
I.
INTRODUCTION
The operation of any computing device is necessarily a physical process, and this funda-
mentally determines the possibilities and limitations of the computing machine. A common
thread in the history of computers is the exploitation and manipulation of different natural
phenomena to obtain newer forms of computing paradigms [1]. For instance, chaos comput-
ing [2], neurobiologically inspired computing, quantum computing[3], and DNA computing[4]
all aim to utilize, at the basic level, some of the computational capabilities inherent in natural
systems. In particular, larger understanding of biological systems has triggered the interest-
ing question: what new directions do bio-systems offer for understanding and implementing
computations?
The broad idea then, is to create machines that benefit from natural phenomena and
utilize patterns inherent in systems to encode inputs and subsequently obtain a desired
output. Different physical principles can yield logic outputs, for logic gates such as AND, OR,
XOR, NOR and NAND which form the basis of the universal general-purpose computation.
It is particularly interesting if systems of biological relevance can also yield logic outputs
consistent with the truth tables of different logic functions [5–7].
In this work we will explore the computational possibilities arising from the dynamics of
a genetic network with quorum sensing feedback. Quorum sensing allows cells to sense their
own density and to communicate with each other, giving the cells the ability to behave as a
population instead of individually. The autoinduction of luminescence in the symbiotic ma-
rine bacterium Vibrio fischeri is perhaps the best characterized example of quorum sensing.
Some other examples include sporulation and fruiting body formation by Myxococcus Xan-
thus and antibiotic production by several species of Streptomyces. The way quorum sensing
is accomplished in cells is through diffusive exchange of auto-inducer molecules which par-
ticipate in intercellular coupling and self feedback. At low cell density, the auto inducer is
synthesized at basal levels and diffuses into the extracellular space and gets diluted. But
as the cell density increases, the intracellular autoinducer concentration increases until it
reaches a threshold beyond which it is produced auto-catalytically. This results in a dra-
matic increase of product concentrations, thus creating bi-stability between a high steady
state and a low steady state [8, 9].
Specifically, in this work we focus on an artificial genetic network known as the repressila-
2
tor, with feedback of the kind that quorum sensing may provide to an isolated repressilator.
A repressilator consists of a ring of three genes, products of which inhibit each other in
cyclic order. The quorum sensing feedback stimulates the activity of a chosen gene provid-
ing competition between the inbitory and the stimulatory activities localized in that gene.
In this work we attempt to use the varying concentrations of proteins in the repressilator,
arising due to quorum sensing feedback, to realize different logic gates.
In the sections below we will first discuss the model of the repressilator genetic network
with quorum sensing feedback, and present the change in dynamics of repressilator proteins
with changing parameters of the quorum sensing feedback. We will then go on to use this
dynamics to implement a range of logic functions, by appropriate input-output associations.
II. REPRESSILATOR GENETIC NETWORK WITH QUORUM SENSING FEED-
BACK
We consider a basic genetic ring oscillator network in which three genes inhibit each
other in unidirectional manner, and an additional quorum sensing feedback loop stimulates
the activity of a chosen gene, providing competition between inhibitory and stimulatory
activities localized in that gene. This system is known to yield behaviour ranging from limit
cycles, to high and low stable steady states [11].
In the model, 3 genes in a cyclic loop forms mRNA (a, b, c) and proteins (A, B, C).
The three genes inhibit each other, where one gene is inhibited by the preceding gene.
Quorum sensing feedback acts between gene B and C, where B stimulates the production
of C by quorum sensing along with an inhibitory action. The model can be reduced to fast
mRNA kinetics, where a, b, c is assumed to be in a steady state with their respective proteins
(A, B, C), leading to da
dt = db
dt = dc
dt ≈ 0.
The equations describing the dynamics of this system then are:
α
α
1 + C n )
1 + An )
1 + Bn +
α
(1)
(2)
(3)
κS
1 + S
)
= β1.(−A +
= β2.(−B +
= β3.(−C +
dA
dt
dB
dt
dC
dt
3
= kS0.S − kS1.B − η.(S − Sext)
dS
dt
(4)
Here S denotes the concentration of the auto-inducer AI.
The relevant parameters in this system are the ratio of protein decay rate to mRNA
decay rate βi, maximum transcription rate in the absence of an inhibitor α, Hill co-efficient
for inhibition n, diffusion co-efficient depending on the permeability of the membrane to
auto-inducer molecule η, the rate of auto-inducer decay to mRNA decay rate kSO, rate of
auto-inducer production kS1, and most importantly, the co-efficient for the production of
C protein due to auto-inducer κ and the concentration of auto-inducer in external medium
Sext responsible for the quorum sensing feedback.
In our simulations we fix the parameters at realistic values [11]: kSO = 1.0, kS1 = 0.025;
η = 1.0; n = 3, and β1 = β2 = β3 = β = 0.1 (relevant to slow protein kinetics). We vary the
parameters that determine the strength of quorum sensing, namely Sext and κ.
On taking different values of parameters Sext and κ, the strength of quorum sensing
changes, and hence the qualitative nature of the dynamics changes. This is clearly evident
from the bifurcation analysis of the concentrations of protein A and B with respect to
parameters Sext and κ displayed in Figs. 1-2, which shows the change in dynamics from limt
cycles to stable steady states as the external concentration of auto-inducer increases. There
is also a sudden switch in the value of the steady state after a critical value of Sext and κ.
III. REALIZATION OF LOGIC GATES
Now, in order to utilize this system to realize a logic gate we need to propose an ap-
propriate input-output mapping, namely, we must first suggest how inputs and outputs are
encoded. The task at hand then, is to ensure that this input and output encoding consis-
tently yields the correct input-output association corresponding to different logic functions,
as displayed in the truth table (Table I).
Specifically, here we will
implement a series of logic gates, such as AND/NAND,
OR/NOR and XOR/XNOR, using the change in dynamics of protein A and B. Further we
will obtain logic operations in parallel by simultaneous measurement of the concentrations
of these two proteins.
4
Encoding Logic Inputs:
Consider the inputs to be encoded by Sext, namely the concentration of auto-inducer in
external medium, which is responsible for the quorum sensing feedback.
We propose that the two inputs I1 + I2 = Sext where I1 and I2 encode the two logic inputs
as follows:
(i) When logic input is 0, I1,2 = 0.0;
(ii) When logic input is 1, I1,2 = 0.2;
So the pair of inputs lead to the following three distinct conditions:
(i) Sext = 0.0
for input set
(ii) Sext = 0.2
for input set
(iii) Sext = 0.4
for input set
(I1, I2) ≡ (0, 0)
(I1, I2) ≡ (0, 1)/(1, 0)
(I1, I2) ≡ (1, 1)
Encoding Logic Output:
We consider the maximum protein concentration as an indicator of the output of the
system. Specifically we consider the maximum concentrations of proteins A and B, denoted
by Amax and Bmax respectively. This gives us the ability to obtain two outputs from the
system in parallel.
Now a prescribed output determination threshold, A∗ and B∗, leads to the mapping of
the concentration to a 0/1 logic output, as folllows:
For protein A:
(i) If Amax < A∗, then the Logic Output is 0
(ii) If Amax ≥ A∗, then the Logic Output is 1
And for protein B:
(i) If Bmax < B∗, then the Logic Output is 0
5
(ii) If Bmax ≥ B∗, then the Logic Output is 1
One can obtain the complementary gates by exchanging the output determination crite-
rion. Namely:
For protein A:
(i) If Amax ≥ A∗, then the Logic Output is 0
(ii) If Amax < A∗, then the Logic Output is 1
And for protein B:
(i) If Bmax ≥ B∗, then the Logic Output is 0
(ii) If Bmax < B∗, then the Logic Output is 1
Controlling the nature of the Logic Function:
We can control the type of logic operation obtained by simply changing the output
determination threshold. Namely, we can obtain AND/NAND, OR/NOR or XOR/XNOR
input-output mappings by considering different A∗ and B∗.
For instance, for realizing the fundamental NAND Logic gate (see Truth table I), we have
the binary logic output determined by the threshold level A∗ = 1. Namely, if the maximum
concentration of protein A, Amax is greater than A∗, then the Logic Output is 1, and if
Amax ≤ A∗, then Logic Output is 0. This is clearly evident in Fig. 3 and Table II.
In order to change the logic function from the fundamental NAND gate to the funade-
mental NOR gate, we simply have to change the prescribed output determination threshold.
Specifically, the system yields NOR logic when the output determination threshold A∗ is 4.
Similarly, using different output determination thresholds B∗ for protein concentration
B, we can obtain AND and XNOR logic operations. This is clearly evident through Fig.
4 and Table III. Again note that the complementary logic functions NAND and XOR are
implemented by a simple toggle of the logic output determination condition.
So we can obtain any combination of AND/NAND/OR/NOR in parallel with
AND/NAND/XOR/XNOR logic through the two protein concentrations for the same pair
6
of inputs. In Fig. 5, the time series plots for Protein A and B exhibits the simultaneous
realization of NAND and XNOR gates respectively, for a representative random stream of
input values. Clearly, this same system can yield XOR and AND in parallel by the alternate
output association, namely a logic output 1 is obtained when the proteins A and B are
below output threshold A∗ and B∗ respectively, and 0 otherwise. In fact, the combination
of AND and XOR in parallel is particularly useful, as it forms the basis of the ubiquitous
bit-by-bit addition.
Alternately, the logic operation may also be changed by a constant shift in input
encoding, namely I1 + I2 = Sext + Slogic, where different values of Slogic yield different logic
functions. For instance, using protein B for output determination, and keeping level B∗
fixed at 15, Slogic = 0 yields XNOR, and Slogic = 0.1 yields NAND. Similarly, using protein A
for output determination, with A∗ = 1, Slogic = 0 yields NAND, and Slogic = 0.2 yields NOR.
Using activation rate parameter κ to encode Logic Inputs:
We now consider an alternate input encoding, with the logic inputs determining κ. Here
Sext is held fixed at 0.0, and varying κ yields different protein concentrations, which leads
to different outputs (see Figs. 6-7).
Specifically, let κ be determined by the logic inputs I1 and I2 as follows:
κ = I1 + I2 + κbase
with I1,2 = 0 when logic input is 0, and I1,2 = 20 when logic input is 1.
Taking a base value of κbase = 10, we have:
(i) κ = 10 when input set (I1, I2) is (0, 0)
(ii) κ = 30 when input set (I1, I2) is (0, 1)/(1, 0)
(iii) κ = 50 when input set (I1, I2) is (1, 1)
The binary Logic Output is obtained from the maximum concentration of proteins A and
B, as before. Namely, we determine the logic output through a threshold: if Amax(Bmax) <
A∗(B∗), then logic output is 0, and if Amax(Bmax) < A∗(B∗), then logic output is 1.
7
As before, the nature of the logic function obtained can be controlled by the output
determination threshold for the proteins A and B. When A∗ = 2 we obtain NAND logic
and for A∗ = 6 we obtain the fundamental NOR operation. Similarly, when B∗ = 40 we
obtain AND and for B∗ = 20 we obtain XNOR. Representative values are displayed in
Tables IV and V.
Hence, XNOR/XOR/AND/NAND and NAND/AND/NOR/OR gates can be realized
in parallel for the same set of inputs. Namely, by setting the appropriate output de-
termination condition, the protein concentrations can yield any of the logic functions
XNOR/XOR/AND/NAND and NAND/AND/NOR/OR independently and simultaneously.
IV. CONCLUSIONS
In summary, we have demonstrated how a genetic ring oscillator network with quorum
sensing feedback can operate as a robust logic gate. We use the concentration of auto-
inducer in external medium Sext, and alternatively an activation rate parameter κ (cf. Eqn.
1), to encode logic inputs. Both these quantities regulate the strength of the quorum sensing
feedback. The concentrations of proteins A and B determine the logic outputs. This input-
output association yields a range of logic functions, namely AND/NAND, OR/NOR and
XOR/XNOR. So the system can act as a versatile unit that can morph between different
logic operations. We further demonstrated the capacity of this system to yield different logic
operations simultaneously through the two proteins A and B.
These observations may provide an understanding of the computational capacity of
systems with quorum sensing feedback. It also may have potential relevance in the design
of biological gates, with the ability to flexibly reconfigure logic operations, and implement
logic operations in parallel. Further, since electronic analogs of such systems have al-
ready been implemented [11], our results may readily lead to bio-inspired computing devices.
[1] J. P. Crutchfield, W. L. Ditto, and S. Sinha, Introduction to Focus Issue:
Intrinsic and
Designed Computation: Information Processing in Dynamical Systems – Beyond the Digital
Hegemony, Chaos 20 037101 (2010)
8
[2] S. Sinha and W.L. Ditto, Phys. Rev. Lett. 81 (1998) 2156; K. Murali, S. Sinha, W.L. Ditto,
A.R. Bulsara, Phys. Rev. Lett. 102 (2009) 104101; W. L. Ditto, A. Miliotis, K. Murali, S.
Sinha, and M. L. Spano, Chaos 20 (2010) 037107
[3] A. Steane (1998) Rep. Prog. Phys. 61 117
[4] DNA Computing: New Computing Paradigms, G. Paun, G. Rozenberg and A. Salomaa,
Springer (Berlin and New York) 1998
[5] Y. Benenson, Science 340 (2013) 554
[6] Ando, H., Sinha, S., Storni, S., and Aihara, K.: Synthetic gene networks as potential exible
parallel logic gates. Europhys. Letts., 93 (2011) 50001.
[7] E. H. Hellen, S. K. Dana, J. Kurths, E. Kehler, S. Sinha, Noise-aided Logic in an Electronic
Analog of Synthetic Genetic Networks, PLoS One (2013) doi:10.1371/journal.pone.0076032
[8] Dockery, J.D., Keener, J.P.:A mathematical model for quorum sensing in Pseudomonas aerug-
inosa. Bull. Math. Biol. 63, 95116 (2001).
[9] S. James, P. Nilsson, G. James, S. Kjelleberg and T. Fagerstrm (2000). Luminescence control
in the marine bacterium Vibrio scheri: An analysis of the dynamics of lux regulation. J. Mol.
Biol. 296, 1127-1137.
[10] Ward, J.P., King, J.R., Koerber,A.J.,Williams, P., Croft, J.M., Sockett, R.E.:Mathematical
model of quorum sensing in bacteria. IMA J. Math. Appl. Med. Biol. 18, 263292 (2001).
[11] Hellen EH, Dana SK, Zhurov B, Volkov E (2013) Electronic Implementation of a Repressilator
with Quorum Sensing Feedback. PLoS ONE 8(5): e62997. doi:10.1371/journal.pone.0062997.
[12] Garcia-Ojalvo, J., Elowitz, M.B. and Strogatz, S.H.: Modeling a synthetic multicellular clock:
Repressilators coupled by quorum sensing. Proceedings of the National Academy of Sciences
of the U.S.A. 101, 10955-10960 (2004).
[13] Potapov, I., Zhurov, B. and Volkov, E.: Quorum sensing generated multistability and chaos
in a synthetic genetic oscillator. Chaos,22, 023117 (2012).
[14] Koseska A, Ullner E, Volkov E, Kurths J, Garc a-Ojalvo J (2010) Cooperative differentiation
through clustering in multicellular populations. J. Theor. Biol. 263: 189202.
[15] Elowitz M, Lim WA (2010) Build life to understand it. Nature 468: 889890.
[16] M.M. Mano, Computer System Architecture, 3rd edition, Prentice Hall, Englewood Cliffs,
1993; Bartee, T.C. Computer Architecture and Logic Design, New York, Mc-Graw Hill, 1991.
9
Inputs (I1, I2) AND OR NAND NOR XOR XNOR
(0,0)
(0,1)/(1,0)
(1,1)
0
0
1
0
1
1
1
1
0
1
0
0
0
1
0
1
0
1
TABLE I: Truth table of the basic logic operations for a pair of inputs: I1, I2 [16]. Note that
NAND and NOR are fundamental logic gates from which any logic circuit can be constructed [16].
Here AND and NAND, OR and NOR, XOR and XNOR are complementary pairs and are simply
given by the NOT operation on the logic output, where NOT gives value 1 when input is 0 and 0
when input is 1.
Inputs
Sext
(0, 0)
(0, 1)/(1, 0)
(1, 1)
0.0
0.2
0.4
Amax NAND Logic with A∗ = 1 NOR Logic with A∗ = 4
≈ 6.0
≈ 2.0
≈ 0.0
1
0
1
0
1
0
TABLE II: The maximum concentration of protein A mirroring the fundamental NAND and
NOR logic gates. The output determination threshold A∗ = 1 for the NAND logic operation and
A∗ = 4 for the NOR logic operation.
Inputs
Sext
(0, 0)
(0, 1)/(1, 0)
(1, 1)
0.0
0.2
0.4
Bmax AND Logic with B∗ = 30 XNOR Logic with B∗ = 15
≈ 20.0
≈ 4.0
≈ 45.0
0
0
1
1
0
1
TABLE III: The maximum concentration of protein B mirroring the AND and XNOR logic gates.
The output determination threshold B∗ = 30 for the AND logic operation and B∗ = 15 for the
XNOR logic operation.
10
Inputs
κ
(0, 0)
10.0
(0, 1)/(1, 0)
(1, 1)
30
50
Amax NAND Logic with A∗ = 2 NOR Logic with A∗ = 6
≈ 9.0
≈ 3.0
≈ 0.0
0
0
1
1
0
1
TABLE IV: The maximum concentration of protein A mirroring the AND and XNOR logic gates.
The output determination threshold A∗ = 2 for the NAND logic operation and A∗ = 6 for the
NOR logic operation.
Bmax AND Logic with B∗ = 40 XNOR Logic with B∗ = 20
Input
(0, 0)
(0, 1)/(1, 0)
(1, 1)
κ
10.0 ≈ 28.0
≈ 5.0
≈ 45.0
30
50
0
0
1
1
0
1
TABLE V: The maximum concentration of protein B mirroring the AND and XNOR logic gates.
The output determination threshold B∗ = 40 for the AND logic operation and B∗ = 20 for the
XNOR logic operation.
11
FIG. 1: Bifurcation diagram displaying the extremal values of the concentration of Protein A
(left) and Protein B (right) vs the auto-inducer concentration in the external medium Sext. The
values of other parameters are: β = 0.1, α = 46 and κ = 14. For every set of input parameters,
the concentrations of proteins are initialised with Aint = Bint = Cint = 0.0.
FIG. 2: Bifurcation diagram displaying the extremal values of the concentration of Protein A (left)
and Protein B (right) vs parameter κ. The values of other parameters are: β = 0.1, α = 46 and
Sext = 0.0. Here the variation of the strength of the quorum sensing feedback solely depends on κ.
For every set of input parameters, the concentrations of proteins are initialised with Aint = Bint =
Cint = 0.0.
12
FIG. 3: Maximum concentration of protein A vs. Sext. The maximum is determined from a range
of around 100 time steps. The fundamental NAND and NOR gates are indicated here, with the
output determination level being A∗ = 1 (shown by green line) for NAND, and A∗ = 4 (shown
by blue line) for NOR logic. The grey encircled values are the logic inputs, with the first circle
at Sext = 0 denoting the input set (0, 0), the second circle at Sext = 0.2 denoting the input sets
(1, 0)/(0, 1), and the third circle at Sext = 0.4 denoting the input set (1, 1) (see text for input
encoding method). Other parameters are: kSO = 1.0; kS1 = 0.025; η = 1.0; n = 3, β = 0.1, α = 46
and κ = 14.
13
FIG. 4: Maximum concentration of protein B vs. Sext. The maximum is determined from a
range of around 100 time steps. The AND and XNOR gates are indicated here, using output
determination threshold B∗ = 30 and B∗ = 15 respectively. Again, the grey encircled values are
the logic inputs, with the first circle at Sext = 0 denoting the input set (0, 0), the second circle
at Sext = 0.2 denoting the input sets (1, 0)/(0, 1), and the third circle at Sext = 0.4 denoting the
input set (1, 1) (same input encoding method as in Fig. 3). Other parameters are: kSO = 1.0;
kS1 = 0.025; η = 1.0; n = 3, β = 0.1, α = 46 and κ = 14.
14
FIG. 5: Time evolution of the concentrations of Protein A (left) and Protein B (right). Here the
output determination threshold (shown as a yellow line), which serves as logic selector, is set at
A∗ = 0.75 and B∗ = 15. The timing sequence in blue (top) shows a random stream of binary logic
input values I1 + I2. The sequence in green (middle) displays the binary logic output desired for
the implementation of NAND (left) and XNOR (right) gates. Note that every time a new input
set is taken, the transient dynamics (≈ 100), indicative of the latency, is not displayed.
15
FIG. 6: Maximum concentration of protein A vs. κ. The maximum is determined from a range of
around 100 time steps. The fundamental NAND and NOR gates are indicated here, using output
determination threshold A∗ = 2 and A∗ = 6 respectively. The grey encircled values are the logic
inputs, with the first circle at κ = 10 denoting the input set (0, 0), the second circle at κ = 30
denoting the input sets (1, 0)/(0, 1), and the third circle at κ = 50 denoting the input set (1, 1) (see
text for input encoding method). Other parameters are: kSO = 1.0; kS1 = 0.025; η = 1.0; n = 3,
β = 0.1, α = 46 and Sext = 0.
16
FIG. 7: Maximum concentration of protein B vs. κ. The maximum is determined from a range of
around 100 time steps. The AND and XNOR gates are indicated here, using output determination
threshold B∗ = 40 and B∗ = 20 respectively. The grey encircled values are the logic inputs, with
the first circle at κ = 10 denoting the input set (0, 0), the second circle at κ = 30 denoting the
input sets (1, 0)/(0, 1), and the third circle at κ = 50 denoting the input set (1, 1) (see text for
input encoding method). Other parameters are: kSO = 1.0; kS1 = 0.025; η = 1.0; n = 3, β = 0.1,
α = 46 and Sext = 0.
17
|
1207.1516 | 1 | 1207 | 2012-07-06T04:19:53 | Mechanical Instabilities of Biological Tubes | [
"physics.bio-ph",
"q-bio.TO"
] | We study theoretically the shapes of biological tubes affected by various pathologies. When epithelial cells grow at an uncontrolled rate, the negative tension produced by their division provokes a buckling instability. Several shapes are investigated : varicose, enlarged, sinusoidal or sausage-like, all of which are found in pathologies of tracheal, renal tubes or arteries. The final shape depends crucially on the mechanical parameters of the tissues : Young modulus, wall-to-lumen ratio, homeostatic pressure. We argue that since tissues must be in quasistatic mechanical equilibrium, abnormal shapes convey information as to what causes the pathology. We calculate a phase diagram of tubular instabilities which could be a helpful guide for investigating the underlying genetic regulation. | physics.bio-ph | physics | Mechanical instabilities of biological tubes.
Edouard Hannezo1, Jacques Prost1,2, Jean-Fran¸cois Joanny1
1Physicochimie Curie (Institut Curie / CNRS-UMR168 /UPMC), Institut Curie,
Centre de Recherche, 26 rue d'Ulm 75248 Paris Cedex 05 France and
2E.S.P.C.I, 10 rue Vauquelin, 75231 Paris Cedex 05, France
(Dated: June 27, 2018)
We study theoretically the shapes of biological tubes affected by various pathologies. When ep-
ithelial cells grow at an uncontrolled rate, the negative tension produced by their division provokes a
buckling instability. Several shapes are investigated : varicose, enlarged, sinusoidal or sausage-like,
all of which are found in pathologies of tracheal, renal tubes or arteries. The final shape depends
crucially on the mechanical parameters of the tissues : Young modulus, wall-to-lumen ratio, homeo-
static pressure. We argue that since tissues must be in quasistatic mechanical equilibrium, abnormal
shapes convey information as to what causes the pathology. We calculate a phase diagram of tubular
instabilities which could be a helpful guide for investigating the underlying genetic regulation.
PACS numbers: xx
Tubular structures are found ubiquitously in living or-
ganisms, from worms to humans and are fundamental
structures of many organs: they convey liquids, gases or
cells throughout the body. Typical examples are arteries,
intestinal tubes or renal excretory canals. The disruption
of these tubes is at the origin of many pathologies, and al-
though a considerable amount of research has focused on
the molecular mechanisms underlying each disease and
each type of tube [1–5], little attention has been given
to the biomechanical aspects of tube stability. Here, we
present a general framework to describe the instabilities
of cellular tubes; we also discuss the possibility to in-
fer the underlying causes of pathologies by studying the
shapes of the abnormally deformed tubes. Indeed, tube
instabilities are observed in many different organs but al-
ways seem to fall into a limited number of physiological
categories : enlarged, sinusoidal, or varicose tubes[2][6].
From a physicist's point of view, these shapes are remi-
niscent of classical fluid instabilities. Nevertheless, a key
difference from passive systems is that pathological in-
stabilities result from the internal activity of the tube.
Cell division, or active fluid pumping, are often the mo-
tor of the instability. Thus, a mechanical description of
the tube shapes could bear a lot of information on the
underlying mechanisms of tube formation and instability.
In the past years, progress has been made in the me-
chanical description of tissues as dividing elastic media[7–
9]. Stresses inside the growing tissues are essential to de-
termine the final architecture of an organ and conversely,
there are feedback mechanisms of form on growth. Here,
we apply these models, which have been usually used
for bulk tissues, to cylindrical geometries. We consider
a wide variety of epithelial tubes that all share key fea-
tures. The tube is composed of a layer of dividing ep-
ithelial cells, which exert pressure on the surrounding
medium, and of an elastic basement membrane which
provides mechanical stability. A softer, visco-elastic con-
nective tissue, which we consider infinite, is surrounding
the tube.
This simple description suggests a competition be-
tween several forces. The fluid-tube interface has a ten-
sion and can undergo a Rayleigh-Plateau like instability
[10], reminiscent of a cylindrical jet breaking into drops.
However, the tube is elastic, and can resist this defor-
mation as discussed for passive lipid tubes [11]. Here we
wish to add active effects: epithelial cells grow (some-
times at an uncontrolled rate), and the tube can buckle
under the load induced by growth. This can be modeled
by a uniform negative surface tension: cell divisions are
driving an increase of the surface area of the tube. As
noted in [7], sufficient mechanical feedbacks can explain
why growth is spatially uniform. We will make that as-
sumption here, given the regularity of the patterns we
wish to explain. Tubes can also dilate because of an ex-
cess uniform fluid pressure.
FIG. 1. Left: Sketch of our model for biological tubes. A sin-
gle layer of epithelial cells rests on an elastic membrane (blue),
surrounded by soft connective tissue (not drawn). Right: Dif-
ferent tube configurations: a) reference, b) varicose/pearling,
c) enlarged, d) sinusoidal e) sausage-like. b) and e) have the
same symmetry, but correspond to different instabilities, as
described in the text.
We successively investigate the morphologies sketched
on Fig.1, in that order. The Young modulus Et for bi-
ological tubes, such as arteries [12], is of the order of
104 − 106P a, which is stiff compared to the surrounding
2
1
0
2
l
u
J
6
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
6
1
5
1
.
7
0
2
1
:
v
i
X
r
a
r(cid:31)hEpithelial cell wallBasement membranea)e)d)c)b)tissues. The reference state is a tube of infinite length
and radius r0. We first discuss a peristaltic perturba-
tion (varicose, or pearling) which respects the cylindrical
symmetry. The displacements can therefore be written
(ur(z), uz(z)) where z is the coordinate along the tube
and r the radial direction. The forces involved are equiv-
alent to those derived from the following effective energy
[13]:
Eel
2πr0
θ +2νezeθ)dz +
(cid:90) (cid:18)
C − 1
r0
(cid:19)2
(e2
z +e2
(cid:90)
K
2
dz
=
Eth
2
(1)
where ν is the Poisson ratio, and the strains are defined
as
eθ =
ur
r0
+
1
2
(
ur
r0
2
) ; ez = u(cid:48)
z +
(u(cid:48)
r
2 + u(cid:48)
z
2)
1
2
(2)
z
where the prime denotes a derivative with respect to z.
The first term is the stretching elastic energy of the tube.
We use here the classical Foppl-Von Karman approxima-
tion [13], neglecting u(cid:48)
2, but we keep the nonlinear term
( ur
)2, if uniform dilations of the tube are non-negligible.
The second term is the layer bending energy where C is
r0
the local curvature, and K = Eth3
12(1−ν2) the bending mod-
ulus. A key difference with Ref.
[14] is that we do not
assume that the outer surface of the cylinder is fixed. The
case of arteries is complicated by the presence of smooth
muscles and anisotropy but as we wish to capture the es-
sential physics, we restrain ourselves to this simple non-
linear and isotropic elastic theory. Surface and pressure
forces, which drive the deformation, are derived from an
effective energy in which the non-equilibrium aspect is
hidden in the surface tension and excess pressure terms:
(cid:90)
Ea =
(2πγ(r0 + ur(z))(cid:112)1 + u(cid:48)
r(z)2
− π∆P (r0 + ur(z))2)dz
(3)
where γ is the effective surface tension, which contains
a negative contribution due to cell division. As a result,
γ can be positive or negative. ∆P is an excess pressure
inside the tube. We first investigate a deformation of the
tube to a radius ur(z) = r0[R0 − 1 + A cos(kz)] which
includes both a uniform dilation to a radius R0r0 and a
pearling deformation of amplitude Ar0 (cid:28) r0. All lengths
have been normalized by r0 (see SI for details). We first
discuss the effect of negative surface tension, assuming
that there is a small excess pressure, ∆P r0 (cid:28) γ(h/r0)2.
Compared to the classical Rayleigh-Plateau instability,
the wavelength observed in vivo is small, close to the
value of the tube radius [2, 6, 15]. We propose that this
pearling instability is due to a buckling of the tube in-
duced by the homeostatic pressure of the dividing ep-
ithelial cells. This hypothesis has biological grounds for
hepatic arteries, since varicoses are associated with Fi-
bromuscular Dysplasia (FMD)[15–17] or polycystic dis-
orders in renal canals [18].
2
(cid:28) γ
In the following, we assume that the value of the ten-
Eth (cid:28) 1, which seems to be the
sion is such that h2
r2
0
case for reasonable values of the parameters : tubes we
≈ 0.01, and as we will discuss below,
will study have h2
r2
γ
Eth ≈ 0.1 . The minimization of the energy with respect
0
γ
Eth (cid:39) 1. In this limit, there is a
to R0 leads to R0 = 1 +
bifurcation from tubular to pearling states for
(4)
γ > γp = Eth
h√
3r0
h√
1
12r0
1−ν2
) 1
at a wavelength λp = 2πr0(
2 . The finite
elasticity of the surrounding medium does not change
these equations as long as its modulus Es is such that
Esr0 (cid:28) Eth, which is the case here. The Poisson modu-
lus ν only has a weak effect. We have performed a numer-
ical minimization and Fig.2 shows a plot of the resulting
shape of a tube that has undergone a pearling instability.
For large arteries such as the common hepatic artery, the
wall to lumen ratio is approximately h
= 0.1, so that
λp ≈ 1.1r0. Strikingly, clinical studies [16] confirm that
r0
the size of the pattern is proportional to the radius of
the artery, with the exact same coefficient as above, in
contradiction to analyses based on the Rayleigh-Plateau
instability [19].
Away from the bifurcation point, the wavelength varies
only slowly. This variation is compatible with the biolog-
ical variability, and explains why the same range of wave-
lengths is always observed. Since the wavelength is of the
order of the radius, we need not consider instabilities in
the section plane of the tube, studied in Ref.
[20]. For
much thinner tubes, such an analysis would be required.
The negative tension exerted by the cells is linked to the
homeostatic pressure, γ ≈ −Phh [21]. Pressure measure-
ments for tumors or dysplasia often give values in the
range of Ph = 104P a [22]. Taking Et = 105P a, the crit-
ical tension γp ≈ 7.103h is perfectly compatible with a
buckling induced by the dividing cells.
FIG. 2. Comparison between the shapes obtained from buck-
ling theory and in vivo data. A: a common hepatic artery
displaying varicosities [15]. B: a drosophila tracheal tube dis-
playing sinusoidal mutation [23], λ ≈ 5r0.
Polycystic diseases are more complex. The tubular
canals of the kidney loose their stability, giving rise to
ABa variety of patterns : string of small pearls, large con-
volutions, or huge spheric cysts that can invade the en-
tire organ [6] [24]. Three mains factors [18] have been
identified : increase in basement membrane compliance,
uncontrolled division of the epithelial cells forming the
tube wall, and disorders in ions pumps, causing excess
water to be pumped in the canals.
In our formalism,
this relates respectively to a decrease in Et, an increase
in Ph, and an increase in ∆P . The excretory canal of
C. Elegans is often used as a model organism to study
these diseases. Although the tube is different in nature
[25] (a single, elongated cell invaginates to form a tube
which spans its entire length), it can be described by
similar equations : there is a bending energy associated
to membrane deformation, a surrounding elastic medium
prevents stretching, the tube can grow because of exces-
sive lipid transport, analogous to negative surface ten-
sion, and there can exist hydrostatic pressure differences.
Thus, in order to incorporate the effect of pressure dif-
ferences, we now consider the opposite limit of enlarged
tubes, where the pressure dominates over tension effects,
∆P r0 (cid:29) γ(h/r0)2 . The equilibrium radius is such that
the elastic deformation equilibrates the excess pressure.
(cid:18)
(cid:19)1/2
R0 =
1 + 2
∆P
Et
r0
h
(5)
The physical picture of pearling is not modified greatly
by considering the effect of an excess pressure. The
threshold tension remains of the same order, and the
characteristic undulation wavelength is slightly reduced
: λp = λp(∆P = 0)(1 + 5∆P r0/3Eth)−1/4.
We then turn to sinusoidal oscillations of the entire
tube. The classical Euler buckling instability of an elastic
column under compression occurs at a wavelength which
is the system size, in the absence of a surrounding elastic
medium. However, in an elastic medium, buckling occurs
at a finite wavelength. We have shown recently that this
instability can be invoked to describe the morphology of
the intestine [26].
We call Et and Es respectively the Young moduli of the
tube and surrounding elastic medium, keeping in mind
that Et >> Es. We denote the deformation of the tube
by w(z, t) = B cos(kz). The curvature modulus of a hol-
low elastic tube is K = πEtr3
0h, the force exerted by
the growing tube is f = 2πγr0, and the elastic force ex-
erted by the surrounding medium is proportional to the
deformation with a coefficient α. The threshold force of
the sinusoidal instability is f = 2
Kα, thus the critical
tension for buckling is
√
EsEt
γ > γs ≈ 0.8
√
√
r0h
(6)
3
0
We now compare sinusoidal deformations, which oc-
cur for cell tensions larger than γs given by Eq.(6),
with pearling deformations with a threshold γp given by
Eq.(4). The preferred morphology essentially depends on
the value of the parameter r = Esr3
Eth3 . If the substrate is
very soft or the tube very hard, the sinuous morphol-
ogy is expected, since in that case, it requires a smaller
tension. Conversely, if h
is very small, the pearling in-
r0
stability is expected. Physically, this is due to the fact
that a larger tube is harder to bend than a thicker tube
: its curvature modulus varies as the cube of the radius
and as the first power of thickness. For a typical tube,
h
= 0.1, and Et = 105P a, the critical substrate rigidity
r0
separating both morphologies is Es = 102P a. This anal-
ysis is important since a loss of tubular rigidity is one
of the causes of cystic diseases. The situation is much
more complex if the substrate is modeled as viscoelastic.
Over long timescales, sinuous deformations are always
expected, but can be kinetically constrained. A detailed
discussion goes beyond the scope of this paper[27].
In the region where both instabilities can be observed,
it is necessary to compare the energies of the two con-
figurations. We plot in thick green the transition line
between the two instabilities, in the vicinity of the criti-
cal point.
FIG. 3. Stability diagram of all possible configurations of a
tube, depending on the relevant parameters. The vertical axis
is the ratio of the moduli of the surrounding tissue and of the
tube. The horizontal axis is the ratio of the tension, negative
or positive, and the elasticity of the tube. Lines in thin red
represent the transition to an instability. The thick green
line separates stability domains of the varicose and sinuous
patterns.
) 1
The wavelength of the instability is λs = 2πr0( h
4 .
A rough estimate leads to λc ≈ 5 − 10 r0, in agreement
2r0
with observations of degenerate excretory canals in C.
Elegans [6], or of the tracheal system in drosophila ([2]
and Fig.2).
Et
Es
The results are displayed on a stability diagram in
Fig.3. The drosophila tracheal system would be a good
candidate for putting this theory to the test. It has been
widely studied as a model of tube size regulation and
the deletion of several genes has been identified to cause
pathologies. Specifically, mutants lacking the Sinuous,
or Varicose genes respectively display sinuous [28], and
pearling state morphologies [29]. Although we do not
have a good understanding of their roles, it is known that
they are required in cell-cell junctions (associated to the
rigidity of the tube) and that they have a role in control-
ling growth. We now show how our phase diagram could
guide further experiments. The modulus of the normal
tracheal wall was inferred from experiments to be around
≈ 0.2. Although tracheal connective
20kP a[30], and h
r0
tissue has not been probed, usual orders of magnitude are
100P a. In normal organisms, the relevant ratio therefore
is Esr3
Eth3 = 0.6 < 1, so sinusoidal instabilities are expected.
Nevertheless, disturbing junctions lowers the rigidity of
the tube, and a three-fold decrease in Et would cause the
pearling instability to be favored. We therefore suggest
that more emphasis should be put on measuring each
gene influence on the mechanical properties of cellular
tubes.
0
A last situation to address is a tube in the absence
of any dysplasia of the cell wall. The surface tension is
then positive.
In this regime, we expect either a sta-
ble tube, or a classical Rayleigh-Plateau instability of an
elastic cylinder. This is slightly different from the situa-
tion encountered in membrane tubes [11], since the fluid
cylinder is surrounded by an elastic shell.
We study a perturbation ur(z) = r0[(R0 − 1) +
A cos(kz)]. The situation that we wish to discuss does
not involve any cell growth, and the instability thus
occurs on a time scale small enough volume conserva-
tion is a reasonable assumption. This imposes that
0 − 1 = − A2
R2
2 . Minimizing the energy with respect to
A (SI) yields a finite value of A only for a tension larger
than the critical value γr = Eth 1−ν2
1−q2 . Below this critical
value of the tension, the cylinder is stable. The instabil-
ity first occurs for the longest wavelength mode q → 0.
Determining the exact most unstable wavelength requires
a non-linear analysis, which is beyond the scope of this
paper.
Nevertheless, if the size of the perturbation approaches
the radius of the tube, disconnected pearls are obtained
and the most unstable wavelength can be found by study-
ing the dynamics of the perturbation uz. The continu-
ity equation and a Poiseuille approximation for the flow
z P where η
lead at lowest order to:
the fluid viscosity. The hydrostatic pressure P is calcu-
lated from the force-balance equation on the membrane
P ∝ (cid:0)γA(−q2 + 1) − Eth(1 − ν2)A(cid:1) cos(qz). The most
∂uz
∂t = r4
0/(16η)∂2
r = γ−Eth(1−ν2)
2γ
unstable wavenumber is then q2
. This
is an explanation similar to [19] for the sausage-like pat-
tern observed on some arteries. Indeed it has been shown
[31] that a very soft elastic cylinder can undergo a classi-
cal Rayleigh-Plateau instability. Testing the case of hol-
low elastic tubes in the same fashion could be an ex-
perimental confirmation to our theory. As expected, if
4
γ/Eth (cid:29) 1, the most unstable wavevector converges to
the wavevector of the classical Rayleigh-Plateau instabil-
ity qRP = 1√
[10].
2
The main result of this paper is the stability diagram,
sketched on Fig.3 presenting the stabilities of most mor-
phologies found in cellular tubes, depending of their me-
chanical properties. Buckling theory is in good quanti-
tative agreement with the in vivo data. We have here
deliberately used simple models for the mechanics of the
tubes, to emphasize the generality of our results, and the
relevance of a few dimensionless ratios to predict the ob-
served instabilities. Most importantly, we want to stress
the fact that tissues' shapes could turn out to be a mean-
ingful criterion to detect the causes of specific patholo-
gies. For example, in the case of excretory kidney canals,
dysplasia leads to a different instability than fluid accu-
mulation. More work is clearly needed in each specific
case, but we do not expect the basic physics presented in
this paper to be modified.
[1] M. Affolter et al., Dev Cell, Vol. 4, 1118, (2003)
[2] G. J. Beitel et al., Development 127, 3271-3282 (2000)
[3] M.A. Baer et al., Curr Top Dev Biol, Vol. 89 (2009)
[4] D. J. Andrew, A. J. Ewald, Dev Biol 341,3455, (2010)
[5] B. Lubarsky, M. A. Krasnow, Cell, Vol. 112, 19-28 (2003)
[6] M. Buechner et al. Dev Biol 214, 227-241 (1999)
[7] B. Shraiman, P Natl Acad Sci USA, 102:3318-3323 (2005)
[8] T. Mammoto et al., Development 137:1407-1420 (2010)
[9] J. Ranft et al., P Natl Acad Sci USA, 107, 20863 (2010)
[10] Lord Rayleigh, Philos. Mag. 34, 145 (1892).
[11] R Bar-Ziv and E. Moses, Phys. Rev. Lett., 73, 10 (1994)
[12] S. Laurent et al., Arteriosclar Thromb Vasc Bio, 14:1223-
131 (1994)
[13] S. Timoshenko, J.M. Gere, Theory of Elastic Stability,
second ed. MacGraw-Hill, New York (1961).
[14] P. Ciarletta et al, J. Mech. Phys. Solids, 60 525-537
(2012)
[15] B. Peynircioglu and B. E. Cil, Cardiovasc Intervent Ra-
diol, 31:S38-S40 (2008)
[16] P. F. J. New, Am Roentgen Ray Soc, 97-1, 488-499 (1966)
[17] P.-F. Plouin et al, Orphanet Journal of Rare Diseases,
2:28 (2007)
[18] L. W. Welling, Pathogenesis of cysts and cystic kidneys.
In The Cystic Kidney (K. D. Gardner and J. Bernstein,
Eds.), 99-116. Kluwer, Amsterdam, Netherlands (1990)
[19] P. Alstrom et al, Phys. Rev. Lett, 82, 9 (1999)
[20] J. Yin et al., P Natl Acad Sci USA, 105,19132 (2008)
[21] M. Basan et al., HFSP Journal, 3(4):265-272 (2009)
[22] G. Helmlinger et al., Nat Med 3, 177- 82 (2007)
[23] P. Laprise et al., Curr Biol 12; 20(1): 55 (2010)
[24] A. P. Evan, J. A. McAteer, Cyst cells and cyst walls.
In The Cystic Kidney (K. D. Gardner and J. Bernstein,
Eds.), pp. 2142. Kluwer, Amsterdam, Netherlands (1990)
[25] F. K. Nelson et al, J. Ultrastruct. Res. 82, 156171 (1983)
[26] E. Hannezo et al, Phys. Rev. Lett. 107, 078104 (2011)
[27] R. Huang, J. Mech. Phys. Solids, 53, 63-89 (2005)
[28] V. M. Wu et al, J Cell Biol, Vol 164-2 (2004)
[29] V. M. Wu et al, Development 134, 999-1009 (2007)
[30] A.M. Cheshire et al, Dev Dyn 237:1874-2888 (2008)
[31] S. Mora et al, Phys. Rev. Lett. 105, 214301 (2011)
5
|
1501.07010 | 2 | 1501 | 2015-03-18T11:32:46 | Active particles in heterogeneous media display new physics: existence of optimal noise and absence of bands and long-range order | [
"physics.bio-ph",
"cond-mat.soft"
] | We present a detailed study of the large-scale collective properties of self-propelled particles (SPPs) moving in two-dimensional heterogeneous space. The impact of spatial heterogeneities on the ordered, collectively moving phase is investigated. We show that for strong enough spatial heterogeneity, the well-documented high-density, high-ordered propagating bands that emerge in homogeneous space disappear. Moreover, the ordered phase does not exhibit long-range order, as occurs in homogeneous systems, but rather quasi-long range order: i.e. the SPP system becomes disordered in the thermodynamical limit. For finite size systems, we find that there is an optimal noise value that maximizes order. Interestingly, the system becomes disordered in two limits, for high noise values as well as for vanishing noise. This remarkable finding strongly suggests the existence of two critical points, instead of only one, associated to the collective motion transition. Density fluctuations are consistent with these observations, being higher and anomalously strong at the optimal noise, and decreasing and crossing over to normal for high and low noise values. Collective properties are investigated in static as well as dynamic heterogeneous environments, and by changing the symmetry of the velocity alignment mechanism of the SPPs. | physics.bio-ph | physics | EPJ manuscript No.
(will be inserted by the editor)
5
1
0
2
r
a
M
8
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
0
1
0
7
0
.
1
0
5
1
:
v
i
X
r
a
Active particles in heterogeneous media
display new physics
existence of optimal noise and absence of bands and long-range
order
Oleksandr Chepizhko1,2,a and Fernando Peruani1,b
1 Laboratoire J.A. Dieudonn´e, Universit´e de Nice Sophia Antipolis, UMR 7351 CNRS , Parc
Valrose, F-06108 Nice Cedex 02, France
2 Department for Theoretical Physics, Odessa National University, Dvoryanskaya 2, 65026
Odessa, Ukraine
Abstract. We present a detailed study of the large-scale collective prop-
erties of self-propelled particles (SPPs) moving in two-dimensional het-
erogeneous space. The impact of spatial heterogeneities on the ordered,
collectively moving phase is investigated. We show that for strong
enough spatial heterogeneity, the well-documented high-density, high-
ordered propagating bands that emerge in homogeneous space disap-
pear. Moreover, the ordered phase does not exhibit long-range order,
as occurs in homogeneous systems, but rather quasi-long range order:
i.e. the SPP system becomes disordered in the thermodynamical limit.
For finite size systems, we find that there is an optimal noise value that
maximizes order. Interestingly, the system becomes disordered in two
limits, for high noise values as well as for vanishing noise. This remark-
able finding strongly suggests the existence of two critical points, in-
stead of only one, associated to the collective motion transition. Density
fluctuations are consistent with these observations, being higher and
anomalously strong at the optimal noise, and decreasing and crossing
over to normal for high and low noise values. Collective properties are
investigated in static as well as dynamic heterogeneous environments,
and by changing the symmetry of the velocity alignment mechanism of
the SPPs.
1 Introduction
We understand by an active particle, a particle that is able to convert energy into
work to self-propel in a dissipative medium. There are several strategies to achieve self-
propulsion. (i) Particles may possess an energy depot and be equipped with a motor 1,
a scenario representative of many biological systems such as moving bacteria [2,3],
a e-mail: [email protected]
b e-mail: [email protected]
1 Motility assays represent a particular case, where motors are not attached to the moving
particles, while the energy depot could be considered as extended over the space [1]
2
Will be inserted by the editor
insects [4,5], and animals [6,7]. (ii) Particles may be able to rectify an external driving
in order to achieve self-propulsion in a given direction, a situation commonly observed
in non-living, artificial active particles such as vibration-induced self-propelled rods
and discs [8,9,10], light-induced thermophoretic active particles [11,12,13,14], chem-
ically driven particles [15,16,17,18,19,20], and rollers driven by the Quicke rotation
effect [21]. Independently of the strategy exploited by the particles to self-propel, these
systems are intrinsically out of (thermodynamic) equilibrium 2 and even in absence of
particle-particle interactions exhibit a non-trivial behavior. For instance, fluctuations
in the self-propelling mechanism can lead to complex transients in the mean-square
displacement of the particles [22] as well as anomalous velocity distributions [23].
Interestingly, most, if not all, active particle systems found in nature take place, at
all scales, in the heterogeneous media: from bacterial motion in natural habitats [24],
such as the gastrointestinal tract and the soil, among other complex environments, to
the migration of herd of mammals across forests and steppes [7]. Despite this evident
fact, active matter research has focused almost exclusively, at the experimental and
theoretical level, on homogeneous active systems [25,26,27,23,28]. Non-equilibrium,
large-scale properties of active systems such as long-range order in two-dimensions
as Vicsek et al. [29] reported in their pioneering paper, the emergence of high-order,
high-density traveling bands [30,31], and the presence of giant number fluctuations
in ordered phases [26,32,33] are all non-equilibrium features either predicted or dis-
covered in perfectly homogeneous systems. Here we show that most of these non-
equilibrium features are strongly affected by the presence of spatial heterogeneities.
Moreover, we show that these properties vanish in strongly heterogeneous media.
More specifically, we extend previous results [34] on the large-scale collective prop-
erties of interacting self-propelled particles (SPPs) moving at constant speed in an
heterogeneous space. We model the spatial heterogeneity as a random distribution
of undesirable areas or "obstacles" that the SPPs avoid. The degree of heterogeneity
is controlled by the average density ρo of obstacles. We provide numerical evidence
that indicates that at low densities of heterogeneities ρo, the SPPs exhibit, below
a critical noise intensity ηc1, long-range order (LRO). For noise intensities η close
to ηc1, the SPPs self-organize, as in homogeneous space, into high-density traveling
structures called "bands". We find that as ρo is increased, bands become less pro-
nounced to the point that for large enough values of ρo they are no longer observed.
Our results indicate that in strongly heterogeneous media, i.e. large values of ρo, the
large-scale properties of the system are remarkably different from what we know of
the Vicsek model [29] in two-dimensional homogeneous media. For instance, orien-
tational order for η < ηc1 is no longer LRO, but rather quasi-long range (QLRO),
with the system exhibiting the maximum degree of order at an intermediate noise
value ηM , such that 0 < ηM < ηc1. Moreover, we provide solid evidence that the
system becomes disordered as η → 0. The numerical data suggests the existence of
a second critical point ηc2, with 0 < ηc2 < ηc1 below which the system is genuinely
disordered. The disordered phase at low η values is characterized by the presence of
large, dense moving clusters. We show that the particle number statistics is consis-
tent with these observations: giant number fluctuations (GNF) are high 3 near ηM
and decrease as η approaches ηc2, with GNF becoming weaker as ρo is increased to
the point that fluctuations become normal. Finally, we investigate and compare static
and dynamical heterogeneous media, as well as SPPs with ferromagnetic and nematic
velocity alignment. We show that in all cases there exists an optimal noise intensity
2 This is due to the energy consumption involved by the self-propelling mechanism and
the energy dissipated to the medium
3 The associated GNF exponent adopts its maximum value near ηM .
Will be inserted by the editor
3
Fig. 1. Interaction of a self-propelled particle (blue) with an obstacle (green). Angles are
given with respect to horizontal axis that is directed from left to right. The moving direction
given by the angle θ(t) evolves in time from an initial direction at ti to a final direction at
tf final. The temporal evolution of the particle position, with respect to the obstacles, is
given by the angle α(t). The initial value of α at ti and its final value at tf are marked with
dashed lines. Notice, that after a collision, θ(tf ) ≈ α(tf ), the difference ∆ = θ(tf ) − α(tf )
is such that ∆ << 1 for either large values of γo or large values of Ro. In the figure, γo = 1
and Ro = 1.
.
that maximizes the ordering in the systems. The reported results might be of a great
importance for the design and control of active particles systems.
The paper is organized as follows. In Section 2 we introduce a general and simple
model for SPPs in heterogeneous media. The collective motion phase, the associated
kinetic phase transition and ordering properties of the model are studied in Sec. 3.
Anomalous density fluctuations are addressed in Sec. 4. In section 5, we investigate
the collective properties of SPPs in a dynamical heterogeneous environment, and
discuss the effect of the velocity alignment symmetry. We summarize our results in
Sec. 6.
2 Formulation of the model
We consider a system of N self-propelled particles, moving with a constant speed v0
on a two-dimensional heterogeneous space of linear size L with periodic boundary
conditions. We express the equations of motion of the i-th particle as:
xi = voV(θi)
θi = g(xi)
γb
nb(xi) Xxi−xj <Rb
sin [q(θj − θi)]
(1)
(2)
+ h(xi, θi) + ηξi(t) ,
where the dot denotes temporal derivative, xi corresponds to the position of the ith
particle, and θi is an angle associated to its moving direction. Equation (1) indi-
cates simply that the particle moves at speed v0 in direction V(θi) = (cos θi, sin θi).
Equation (2) conveys the dynamics of the moving direction of the particle, which is
parametrized by the angle θi. The first term on the right-hand side corresponds to a
velocity-velocity alignment mechanism acting between neighboring particles as in the
Vicsek model [29], the second term models the interaction of the i-th particle with
4
Will be inserted by the editor
(a)
(b)
(c)
o
r
c
i
M
o
r
c
a
M
Fig. 2. Simulation snapshots of different phases for ρo = 2.55 × 10−3 for a system size
Nb = 19600 (L = 140). The SPP are represented as black arrows, while the obstacles as green
dots. The bottom panels correspond to macroscopic phases, while the top panels to zoom
up regions, where the interaction between the SPPs and the obstacles can be appreciated.
From left to right: (a) clustered phase, η = 0.01 and r = 0.58, (b) homogeneous ordered
phase, η = 0.3 and r = 0.97, and (c) ordered band phase, η = 0.6 and r = 0.73.
the spatial heterogeneities, and the third term is an additive noise, where hξi(t)i = 0,
hξi(t)ξj (t′)i = δijδ(t − t′), and η denotes the noise strength. The velocity-velocity
alignment is characterized by three parameters: its symmetry, given by q and that
we fix for most of this analysis to be ferromagnetic with q = 1, the interaction radius
Rb, and the alignment strength γb, where the symbol nb(xi) represents the number of
SPPs at a distance less than or equal to Rb from xi, i.e. the number of neighbors of the
i-th particle. The function g(xi) controls the relative weight between alignment (to
the other particles) and heterogeneity/obstacle avoidance. We tested two possibilities,
both leading to the same macroscopic behavior: g(xi) = 1 and g(xi) = 1 − Θ(no(xi)),
where Θ(x) is Heaviside step function. The latter implies that in the proximity of an
obstacle, the SP particle focuses on avoiding it, without aligning to the neighbors.
Results shown here correspond to this definition. Finally, the interaction with the
spatial heterogeneities, which we refer to as "obstacles" or undesirable areas, is given
by the function h(xi, θi), defined as:
γo
(3)
h(xi, θi) =
sin(αk,i − θi) ,
no(xi) Xxi−yk<Ro
if no(xi) > 0, otherwise h(xi) = 0. The term no(xi) represents the number obstacles
at a distance less than or equal to Ro. The angle αk,i is simply the angle of the polar
representation of the vector xi − yk = Γk,i (cos αk,i, sin αk,i), where yk is the position
of the kth-obstacle and Γk,i is the norm of the vector. The position yk of obstacles,
with k ∈ [1, No], is random and homogeneously distributed in space. The interaction
between a SPP and an obstacle is depicted in figure 1: as the particle approaches the
obstacle, its trajectory is deflected. While here we focus mainly on "obstacles" whose
position is fixed in time, we will also address briefly, at the end of the paper, the case
of moving obstacle, i.e. free to diffuse around the space.
It is worth analyzing few simple limits of the model given by equations (1) and (2).
For γb = γo = 0, the equations define a system of non-interacting persistent random
Will be inserted by the editor
5
η=
0.1
0.3
0.7
10
5
obstacle configurations
1
r
0.5
0
Fig. 3. Average values of order parameter r for 10 different initial configuration of obstacles
for three values of the noise strength η and ρo = 0.0325. System size Nb = 10000. Notice that
the average value of the order parameter does not depend on the particular configuration of
the randomly placed (static) obstacles. The error bars correspond to the standard deviation
of the time series of r obtained for each obstacle configuration and set of parameters.
walkers characterized by a diffusion coefficient Dx = v2
o/η2. With γo = 0 and γb > 0
(or equivalently γo > 0 and No = 0), the model reduces to a continuous-time ver-
sion [35] of the Vicsek model [29]. For γo > 0, γb = 0, and No > 0, the equations
describe a system of non-interacting active particles moving at constant speed on
an heterogeneous space, where several interesting non-equilibrium features can be
observed [36]. At low obstacle density, particles move diffusively with a diffusion co-
efficient that is, interestingly, a non-monotonic function of the obstacle density. It
reaches a minimum at a given, non-trivial, non-zero obstacle density. On the other
hand, at high obstacle density and for large enough interaction strength γb, sponta-
neous trapping of particles can occur. In this scenario, particles move sub-diffusively
across the space, spending arbitrary long time in the spontaneously formed traps [36].
Here, we focus on the general scenario where γo > 0, γb > 0, and No > 0. We reduce
the parameter space by fixing the following parameters: Rb = Ro = 1, γb = γo = 1,
ρb = Nb/L2 = 1, v0 = 1 and a discretization time step to ∆t = 0.1. Notice that our
main control parameters are the noise intensity η and the number of obstacles No or
equivalently, the obstacles density ρo = No/L2.
3 The order-disorder transition
The system exhibits three distinct macroscopic phases, for any given obstacle density
ρo > 0, as we move from high to low noise amplitude η. At high noise values, particles
are homogeneously distributed in space as a disordered gas of non-interacting parti-
cles. Below a critical noise value ηc1, the system undergoes a kinetic phase transition
from the disordered gas to a (locally) ordered phase. For finite size systems, there
is a symmetry breaking and the ordered phase implies the existence of a net flux of
particles in a given direction, or equivalently the existence of a preferred direction
of motion. For values close to onset of collective motion, i.e. close to ηc1, and for
rather low values of ρo particles self-organize into high-density, high-order traveling
structures called "bands", as illustrated in Fig. 2(c). As η is decreased further, getting
deeper in the ordered phase, bands disappear and we observe an ordered phase where
particles are roughly homogeneously distributed in space, though anomalously large
6
Will be inserted by the editor
(a)
1
r
0.5
1
r
0.1
ρ
ox102=
0
0.254
1.27
3.25
5.10
6.25
7.65
10.2
(b)
η=
0.1
0.2
0.3
0.6
0.7
0.75
0
0
0.4
η
0.8
0.01
0
0.2
ρ
o
0.4
Fig. 4. (a) Order parameter r vs. noise strength η for various values of the obstacle density
ρo. (b) Order parameter r vs. obstacle density ρo for various values of noise strength η.
System size: Nb = 19600 (L = 140). Notice in (a) that curves for ρo > 0 exhibit a (local)
maximum. This implies the existence of optimal noise value that maximizes collective motion.
density fluctuations are present, Fig. 2(b). If η is decreased even further, coming close
to the noiseless limit, counterintuitively the degree of order drops dramatically and
particles are organized into densely packed moving clusters, see Fig. 2(a), that are
only weakly correlated due to the constant deflections they experience when running
into obstacles.
In order to quantify the degree of order in the system, we use the following order
parameter:
1
Nb
Nb
Xi=1
,
(4)
r = hr(t)it =*(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
eiθi(t)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
+t
where h. . . it denotes the average over time 4. The definition of r is simply, in complex
notation, the norm of the average velocity of the particles. Values of r larger than
zero indicate that there is a net flux of particles in a given direction, while r =
0 corresponds to a disordered system. It is important to stress that the average
value of r does not depend on the particular configuration of (static) obstacles, as
long as the obstacles have been placed randomly in the space. For instance, if we
compare simulations performed with different obstacle configurations, we obtain the
same average value of r, as shown in Fig. 3, where the error bars represent the
measured standard deviation in the corresponding time series of the order parameter
r. In summary, the value of r, Eq. 4, depends only on the noise amplitude η and
obstacle density ρo.
Fig. 4 shows the dependency of the order parameter r with respect to the noise
intensity η for various obstacle densities ρo, panel (a), and with respect to the obstacle
density for various noise intensities, panel (b). The curve that corresponds to ρo = 0 in
Fig. 4(a), black curve, exhibits the behavior that we expect in an homogeneous space,
i.e. as expected in the classical formulation of the Vicsek model [27]. This reference
curve indicates that in an homogeneous system, the maximum order is reached in
the noiseless limit. Notice that for an homogeneous system, as the noise intensity
is increased, the order parameter r monotonically decreases until the critical noise
strength ηc1, above which the system is fully disordered and r = 0. Fig. 4(a) shows
that the presence of even a small amount of obstacles leads to a qualitatively different
picture, with the order parameter r exhibiting a different behavior with η. All curves
with ρo > 0 reach a maximum at a non-zero value of η, and all decrease as the noiseless
4 Since our initial condition is typically a random distribution of particles over space with
random moving direction, this average is performed after removing the transient.
Will be inserted by the editor
7
(a)
0.2
χ
(b)
ρ
o=2.55x10-3
η =
0.802
0.801
0.800
2000
)
r
(
n
ρ
o=2.55x10-3
0.4
η
η
c1
0.8
0
0
(c)
0.15
r
0
0.3
η =
0.802
0.801
0.800
0.6
G
0
0
0.3
r
0
0
2×105
t
4×105
Fig. 5. Evidence of a first-order phase transition at low obstacle densities. (a) Binder cumu-
lant G and susceptibility χ as function of the noise intensity η. Notice that around η ∼ 0.8,
G reaches negative values and χ peaks. (b) Histogram n(r) of the order parameter obtained
from time series of r. (c) Time series of the order parameter r. Notice the flip-flops in the
time series. System size Nb = 19600 (L = 140).
limit is approached. This means that for each value of ρo, there is an optimal value
of η that maximizes collective motion. We refer to this η value as ηM . On the other
hand, we observe that if we fix the noise intensity η and vary the obstacle density
ρo, as displayed in Fig. 4(b), the order parameter r monotonically decreases as ρo is
increased. Notice that curves corresponding to different noises exhibit a quite distinct
behavior, for instance compare the curve for η = 0.3, close to the optimal noise ηM
for most ρo values, with the other curves.
In the following we divide our quantitative analysis into two statical data sets
that correspond to low obstacle density and high obstacle density, respectively. We
show that at quite small obstacle densities, the numerical data is consistent with a
discontinuous (kinetic) phase transition. The numerical data indicates that as the
obstacle density increases, the traveling bands become weaker until they disappear.
We show that once we reach such obstacle densities (i.e. for ρo ≥ 0.1), the order
is no longer long-range (LRO) in the ordered phase, but rather quasi-long range
(QLRO). Our results unambiguously indicate that at such high obstacle densities, as
we approach the noiseless limit, i.e. when particles self-organize into densely packed
moving clusters as shown in Fig. 2(a), the system is fully disordered, which suggests
the existence of a second critical point.
3.1 Low obstacle density
To characterize the phase transition to orientational order at low obstacle densities, we
introduce two additional quantities, the susceptibility χ and the Binder cumulant [37]
8
1
r
0.8
Will be inserted by the editor
(a)
η =
0.05
0.1
0.3
0.4
0.6
1
r
0.1
103
104
Nb
105
0.01
102
(b)
η =
0.05
0.1
0.4
0.6
0.7
103
104
Nb
105
(c)
η
c1
D
0.5
µ
O
R
L
Q
η
c2
0
0
η
0.5
D
1/16
1
Fig. 6. Finite size analysis at low and high obstacle densities. (a) Scaling of r as func-
tion of the system size Nb for low obstacle densities, here ρo = 2.55 × 10−3, for vari-
ous values of η (color coded). The solid lines correspond to exponential fittings: r(Nb) =
r∞ + C∗ exp(−Nb/N∗), with r∞, C∗, and N∗ fitting constants. Notice that for Nb → ∞, the
order parameter reaches a constant value r(Nb → ∞) → r∞, i.e. long-range order (LRO).
(b) Scaling of r as function of the system size Nb for high obstacle densities, here ρ = 0.102,
for various values of η (color coded). The solid curves correspond to power-law fittings:
r(Nb) = A N −µ
, where A and µ are fitting constants. We define, see text, quasi-long-range
order (QLRO) when µ < 1/16. On the other hand, µ = 1/2 implies that the system is fully
disordered. (c) Exponent µ as function of the noise intensity η for ρ = 0.102. The diagram
allows us to define two critical points, indicated by the vertical lines: the vertical line to the
left corresponds to ηc2, while the other one to ηc1. In between ηc2 < η < ηc1 the order is
QLRO. The value 1/16 is indicated by an horizontal red line.
b
G, whose definitions are given by:
χ = hr2it − hri2
t ,
G = 1 −
hr4it
3hr2i2
t
, .
(5)
(6)
We use χ to determine precisely the position of the critical point ηc1 and G to estimate
whether the probability distribution of the order parameter r is unimodal and Gaus-
sian. Notice that the definition of G is directly related to the excess kurtosis. Fig 5(a)
shows both quantities as function of the strength η for one of the smallest obstacle
densities tested, ρo = 2.54 × 10−3. The peak of χ and the sudden change of behavior
of G at η = 0.8 indicates that the critical point is located at ηc1 = 0.8. The drop of G
to negative values suggests that the probability distribution of r is bimodal [38], as
confirmed in Fig. 5(b). This finding is the result of abrupt transitions in the value of
r along time, often called "flip-flops", from low values (disordered gas phase) to high
values (ordered phase) and vice versa, Fig. 5(c). In summary, the statical data close
to the critical point ηc1 is consistent with a discontinuous (kinetic) phase transition:
negative values of G, a bimodal distribution, and flip-flops as η → ηc1.
Our next step is to determine whether at low obstacle densities the observed order
for η < ηc1 remains present in the thermodynamical limit. This involves a finite size
scaling. The goal is to obtain the scaling of the order parameter r with the system
size. If the system exhibits long-range order (LRO), by increasing the system size,
while keeping constant the particle density ρb and obstacle density ρo, we expect r
5. Fig. 6(a)
to saturate to a non-zero value. As measure of the system size we use Nb
shows that at low obstacles densities, here ρo = 2.55 × 10−3, r effectively saturates
5 As measure of system size, instead of Nb, we can use either No or L. Since ρb and ρo are
constant, knowing either L, Nb, or No, we can determine the other two.
Will be inserted by the editor
9
0.03
)
x
(
L
ρ
0.015
moving
direction
(a)
ρο=
0.00765
0.0127
0.0255
0.04
0.1ρ
h
(b)
3
1
-
h
ρ
/
x
a
m
L
ρ
1
0
0
50
x
100
0.01
ρ
o
0.1
Fig. 7. Bands. (a) The particle density profile ρL(x) of the bands along the (band) moving
direction. The horizontal line indicates the density profile corresponding to an homogeneous
distribution of particles whose value we denote by ρh. (b) Maximum ρL max of the density
profile ρ(x)L vs. the density of obstacles. The panel shows the quantity ρL max/ρH −1, which
drops to zero for large densities, which indicates the absence of bands in the system.
with Nb to a non-zero value for a large range of η < ηc1 values 6. Thus, the numerical
data at very low densities, up to the system size we could reach, is consistent with
LRO. This implies that in the thermodynamical limit we expect the system to remain
ordered, i.e. as Nb → ∞, r → r∞(η) > 0, where r∞(η) is the asymptotic value of r
in an infinite system, which is a function of η, ρb, and ρo.
At low obstacle densities, as mentioned above, we observe close to the critical
point ηc1 a behavior consistent with a first-order (discontinuous) phase transition.
This seems to be related [30] with the emergence of high-order traveling bands 7,
as shown in Fig. 2(c). Bands are narrow structures that expand through the whole
system, elongated in the direction perpendicular to their direction of motion. Several
density profiles, corresponding to various ρo values, are displayed in Fig. 7(a). To
construct these profiles, one needs to project the particle positions onto the moving
direction of the band and make a histogram. Bands exhibit a sharp front and a smooth
tail. They move across a background gas of disordered (or weakly ordered) particles
and accumulate particles in the front as they advance forward, while loosing particles
in the rear. The presence of obstacles strongly affects bands. As ρo is increased,
profiles get smoother as illustrated in Fig. 7(a). This can be more quantitatively
seen in Fig. 7(b) that shows how the bands vanish for large value of ρo. In short,
as the number of obstacles is increased, bands become weaker, with a profile that
relaxes towards the background gas. At some point, bands and the background gas
are undistinguishable and bands vanish. This is evident for ρo > 0.1, where bands are
no longer observed.
3.2 High obstacle density
At high enough obstacle densities, where bands are no longer observed, the system
behavior is remarkably different. The finite size study reveals that the system is
unable to reach LRO at such high ρo values. We find that the order parameter r
decays with the system size Nb as a power-law: r ∝ N −µ(η)
, where the exponent
b
6 Unfortunately, we can not be sure that this behavior remains for η < 0.05. Smaller noises
are hard to explore numerically.
7 For instance, when flip-flops are observed, high values of r coincide with the emergence
of bands. Along the simulation, bands appear and disappear constantly, as flip-flips do.
10
Will be inserted by the editor
σ(l)
103
102
101
β
0.7
0.65
0
(a)
(b)
ρ
o=
0.00255
0.153
0.6
η =
0.01
0.3
1.1
102
(c)
<n>l
104
102
(d)
<n>l
QLRO
η
0.3
e
s
a
h
p
r
e
d
r
o
0.6
0.2
0.4
ρ
o
0.6
σ(l)
103
102
101
104
0.8
β
0.7
0.6
Fig. 8. Number fluctuation statistics. (a) Scaling of the variance of the particle number σ(l)
vs. the average particle number hnil for different values of noise strength η for ρo = 0.102.
(b) Scaling of σ(l) vs hnil for different values of the density of obstacles ρo and fixed noise
strength η = 0.3. In (a) and (b), the scaling σ(l) ∝ hni1
- used as reference
- are indicated by a solid red and a dashed blue curve, respectively. The dashed lines in
(a) and (b) represent power-law fitting curves. (c) The dependency of the scaling exponent
β (see text for more details) on the noise strength η at fixed ρo = 0.102. For ρo = 0.102,
ηc1 ≈ 0.45 ± 0.05 and ηc2 ≈ 0.15 ± 0.05, indicated by two dashed vertical lines in the
figure. In the range ηc2 < η < ηc1, the system exhibits Quasi-Long-Ranged Order (QLRO).
(d) Dependency of the scaling exponent β on the density of obstacles at fixed noise strength
η = 0.3. The region where the system exhibits ordered phases is indicated by a vertical
dashed line, where order is first LRO and then QLRO. System size Nb = 40000 (L = 200).
l and σ(l) ∝ hni1/2
l
µ(η) is a function of η, as shown in Fig. 6(b) for ρo = 0.102. For noises close to
the optimal value ηM , we obtain exponent values such that 0 < µ < 1/16. Here, by
analogy with Kosterlitz-Thouless transition [39] we say that when 0 < µ < 1/16 there
is quasi-long range order (QLRO), while for µ > 1/16 we assume that the system is
disordered. Fig. 6(c) shows that the behavior of µ(η) is such that we can define two
disordered phases, one at high noise values and the other one at low noises, and the
ordered phase with QLRO at intermediate noises. This implies the existence of two
critical points, which obey µ(ηci) = 1/16, with i = 1, 2 such that ηc2 < ηc1, see
Fig. 6(c). The system exhibits QLRO when ηc2 < η < ηc1, while being disordered
for η < ηc2 and η > ηc1. Notice that for both disordered phases, µ reaches 1/2, in
particular we observe that µ → 1/2 in two limits: η → 0 and η → ∞. A scaling
r ∝ N −1/2
corresponds to a fully disordered system with a random distribution of
moving directions. Interestingly, the densely packed moving cluster phase at low η
values, as the one observed in Fig. 2(a), corresponds, at high obstacle densities (i.e.
for ρo ≥ 1), to a fully disordered phase. The presence of only QLRO, or rather the
absence of LRO, implies that in the thermodynamical limit we expect r → 0 for all
η > 0 values, i.e. we expect an infinite system to be disordered. Nevertheless, between
ηc2 < η < ηc1, we expect the SPPs display large correlations in their moving direction
also in the thermodynamical limit.
b
Will be inserted by the editor
11
4 Anomalous density fluctuations and clustering statistics
A way to characterize the distribution of SPPs over the space is through the study
of density fluctuations. Particularly useful information is provided by the so-called
number fluctuations. The idea is to divide the space over which particles move in
cells of linear size l and count the number of particles in each cell. Let us call n(xi, l)
the number of particles in the cell of linear size l whose center is at position xi. We
are interested in computing the average of this quantity hn(xi, l)ii and its standard
deviation σ. It can be shown [26] that hn(xi, l)ii = ρbl2 = hnil and that
σ(l) =qh(n(xi, l) − hnil))2ii = hniβ
l ,
(7)
where the average h. . .ii is performed over the cells the space has been divided into.
An exponent β = 1/2 is expected for a random distribution or particles, while β >
1/2 corresponds to giant number fluctuations (GNF). It has been predicted that in
spatially homogeneous systems, the SPPs in the ordered phase -- far away from the
band regime -- exhibit GNF [28]. This prediction has been observed in simulations of
SPP particles, where a value of β ∼ 0.8 was found [38,40]. Here, we are interested
in knowing the impact of an heterogeneous environment on the number fluctuations.
Figure 8 shows how number fluctuations are affected by changing the noise intensity
η for a fixed density of obstacles ρo ≈ 0.102 (see Fig. 8(a),(c)) and by varying the
obstacle density ρo while keeping the noise fixed, here η = 0.3 (see Fig. 8(b),(d)).
The top panels of Fig. 8 correspond to the scaling of σ(l) with the average number
hnil of particles per cell. In the figure, dashed blue lines correspond to a slope 1/2,
while red solid lines to a slope 1. Undoubtedly, GNF are also present in heterogeneous
space. Nevertheless, there are important differences with what we know from SPP in
homogeneous space. For instance, for a fixed (high enough) obstacle density ρo, GNF
are suppressed, or at least decrease, as η → ηc2, i.e. β adopts smaller values as η
approaches ηc2, Fig. 8(c). We recall that in homogeneous space GNF are expected to
be characterized by the same anomalous exponent as η → 0. We also point out that
in both homogeneous and heterogeneous space, number fluctuations become normal
for η > ηc1, i.e. in high-noise disordered phase. This means that at high obstacle
densities (i.e. for ρo ≥ 0.1), GNF are stronger -- meaning that β adopts its largest
value -- at some point in between ηc2 < ηM < ηc1, and this seems to occur close to
ηM . On the other hand, if we fix the noise intensity η, we observe that GNF decay
(i.e., β goes down) as the density of obstacles ρo is increased, reflecting the global
tendency of the system to go to disorder when ρo → ∞. We find that for ρo → 0,
β → 0.8 as expected in an homogeneous media, while as ρo is increased, the exponent
β exhibits two regimes with ρo, approaching linearly for high obstacle densities 1/2,
where fluctuations can be considered normal and the system disordered, see Fig. 8(d).
Another alternative to study how particles are distributed in space is to look
at the cluster size distribution. As before, we are interested in understanding how
the presence of obstacles affects the non-equilibrium clustering statistics of the SPPs
with respect to what we know from homogeneous media [41,42,43]. By "cluster"
we understand a group of connected particles, such that the distance between two
connected particles is smaller or equal to the interaction radius. The size or mass of
a cluster, which we denote here with the letter "m", is the number of particles the
cluster contains. Our quantity of interest is the (weighted) cluster size distribution
(CSD) P (m). Its definition is given by:
P (m) = lim
t→∞
P (m, t) = lim
t→∞
m nm(t)
N
,
(8)
where nm(t) refers to the number of clusters of mass m that are present in the
system at time t. The limit is to indicate that we look at the steady state CSD and
12
Will be inserted by the editor
P(m)
10-2
10-3
10-4
(a)
~m-0.5
~m-1.18
ρ
o=3.25x10-2
η =
0.3
0.6
0.9
(b)
ρ
o =
2.5x10-3
0.4
~m-1.18
η=0.6
P(m)
10-2
10-3
m
102
100
102 m
Fig. 9. Clustering statistics. (a) The cluster size distribution (CSD) P (m) for the fixed
density of obstacles ρo = 3.25 × 10−2 for different noises η. The critical noise values are
ηc2 ≈ 0.05 and ηc1 ≈ 0.75. The CSD for η = 0.9 corresponds to a disordered phase, while
the other two CSDs to ordered phases. The dashed and dotted-dashed curves provide the
scaling ∝ m−1.18 and ∝ m−0.5 used as reference. (b) P (m) for fixed noise value η = 0.6 and
two different values of the obstacle density ρo. With ρo = 2.5 × 10−3 the system is in an
ordered phase and bands are observed, while with ρo = 0.4 the system is fully disordered
and bands are not observed. The dashed line corresponds to the scaling ∝ m−1.18. System
size Nb = 19600 (L = 140).
neglect transitory behaviors. Fig. 9(a) shows how the CSD is changed by varying the
noise η for fixed ρo = 3.25 × 10−2. As a reference, the CSD corresponding to a fully
disordered phase, i.e. η = 0.9, is shown. We find that in between ηc1 and ηM the CSD
distribution is roughly power-law, P (m) ∝ m−ω, with an exponent ω ∼ 1.18 that falls
in the range [0.8, 4/3] as expected [43]. As we move to lower noise values, e.g. η = 0.3,
there is a strong depletion of isolated particles and small clusters and particles tend to
form larger clusters, way larger than those observed close to ηc1 or in the disordered
phase (notice the log scale in the figure). Fig. 9(b) displays the CSD at fixed noise
η = 0.6 and two different values of ρo. For ρo = 2.5 × 10−3 we observe bands and the
CSD is again a power-law with exponent ω ∼ 1.18, similar to what was reported for
the Vicsek model [29] in the band regime where ω ∼ 1.3 [43]. The figure evidences
that by increasing the density of obstacles ρo, at fixed noise, the functional form
of CSD is dramatically affected. In particular, it shows that at very large obstacle
density the CSD becomes exponential as expected for the disordered phase, with a
well defined average cluster size 8 and not surprisingly the system exhibits normal
number fluctuations.
5 Discussion: static vs dynamic heterogeneities and the symmetry
of the interactions
There is no reason to believe that static and dynamic heterogeneous environments
lead to similar large-scale collective effects. In particular, the conclusions drawn from
the finite size analysis performed with static obstacles, i.e. with a sort of "quenched"
noise, may not apply to dynamical heterogeneities. One may argue that dynamical
heterogeneities may be mapped to an effective noise in a SPP system with homoge-
neous space. In this scenario, the dynamical heterogeneities should not affect qualita-
tively the large-scale properties of the system but only have an impact on the critical
point. A rigorous analysis would require a finite size study of SPPs in dynamical het-
erogeneous environments, which is a very time-demanding numerical task out of the
8 Power-law CSDs may be such that their first moment diverges, while exponential CSDs
always have a well defined first moment.
Will be inserted by the editor
13
Fig. 10. SPP in a dynamical environment, where obstacle diffuse with a diffusion coefficient
Do. (a) Order parameter r vs. η for various obstacles densities ρo and constant diffusion
coefficient Do. (b) r vs. η for constant density of obstacles ρo and various diffusion coefficients
Do. Notice that there exists an optimal noise value even in a dynamical environment. System
size Nb = 10000 (L = 100).
scope of the current paper. Less ambitious but not less informative, we can analyze
the impact of a dynamical heterogeneous medium on the collective properties of SPPs
in a fixed system size. Let us assume that the obstacles now diffusive over the space
with a diffusion constant Do. The position of the k-th obstacle obeys:
yk =p2Doξk(t)
(9)
where hξk(t)i = 0 and hξk(t)ξf (t′)i = δ(t − t′)δk,f . Fig. 10 shows the order parameter
r as function of the (angular) noise η, for various values of ρo and fixed obstacle
diffusion coefficient Do = 0.7, panel (a), and fixed obstacle density ρo = 0.102 and
various obstacle diffusion coefficient Do, panel (b). We find that even for a dynam-
ical heterogeneous environment, there is an optimal noise that maximizes collective
motion. As the obstacle density is increased, the level of ordering, i.e. r, decreases
for all angular noises, Fig. 10(a). On the other hand, we learn that the faster the
obstacles diffuse, the weaker is the effect of the obstacles, Fig. 10(b). Moreover, the
numerical data suggests that in the limit of Do → ∞ the system behaves again as
an homogeneous system with its critical point shifted to smaller noise values 9. This
suggests that in this limit effectively the problem can be mapped to an homogeneous
system with an effective (angular) noise intensity.
replace the order parameter by: S2 = D(cid:12)(cid:12)(cid:12)
Finally, we may wonder whether the observed optimal value is due to the par-
ticular symmetry of the velocity alignment between the SPPs that has been used,
i.e. due to q = 1 in Eq. 2. To address this question, we perform simulations with
SPPs interacting via a nematic velocity alignment, i.e. q = 2 in Eq. 2, that move in
an heterogeneous medium with static obstacles. Since the alignment is nematic, we
, where h. . .it represents
a temporal average after a short transient. For a totally disordered system, S2 = 0,
while S2 > 0 implies that the system exhibits, for the tested system size, (global)
nematic order. Fig. 11 shows that optimal noise ηM exists also for SPPs interacting
via nematic alignment in an heterogeneous environment, as discussed above for fer-
romagnetic velocity alignment. In this case, the optimal noise ηM maximizes nematic
ordering, i.e. it favors the emergence of a preferred direction of motion, where 50%
of the SPPs move roughly parallel to it and the other 50% antiparallel to it. Details
about SPPs with nematic velocity alignment in homogeneous media can be found
in [35,44,45].
1
Nb PNb
i=1 ei 2 θi(t)(cid:12)(cid:12)(cid:12)Et
9 Compare Fig. 10(b) and Fig. 4(a), curve for ρo = 0 to see the shift in the critical point.
14
Will be inserted by the editor
1
0.8
0.6
0.4
0.2
2
S
ρ
o=
0
0.001
0.005
0.01
0
0
0.1
0.2
0.3
0.4
η
0.5
0.6
Fig. 11. SPPs with nematic alignment in a (static) heterogeneous medium. Nematic order
parameter S2 vs. noise intensity η for different obstacle densities ρo. Notice that also SPPs
with nematic alignment exhibit an optimal noise that maximizes order in the system. System
size Nb = 10000 (L = 100).
6 Conclusions
We have learned that even small levels of heterogeneity lead to qualitative changes
in the large-scale properties of SPP systems interacting via a velocity alignment
mechanism. Some of the new statistical features that emerge due to the presence
of heterogeneities -- as for instance the existence of an optimal noise that maximizes
order [34] -- are present in both statical and dynamical heterogeneous media, as well
as by changing the symmetry of the velocity alignment mechanism of the SPPs. Other
findings, as the absence of long-range order for high levels of heterogeneity (i.e. for
ρo ≥ 0.1) apply exclusively to static obstacles. In general, we can conclude that in
heterogeneous environments the physics of SPP systems is different from what we
know from homogeneous ones, with the presence of obstacles making more difficult
for the SPPs to spread information about their moving direction across the system.
From the two information spreading mechanisms [48,49] -- the one involving direct
particle-particle interaction and responsible of order inside individual clusters, and
the other one involving cluster-cluster information exchange, often through particle
exchange among clusters -- obstacles affect the second one 10. In particular, due to the
obstacle presence, clusters get quickly uncorrelated as result of independent collisions
with the obstacles. At high obstacle densities or low noise amplitudes, particle ex-
change among clusters becomes insufficient to maintain the moving clusters correlated
and the level of (global) order decreases.
The spatial arrangement of particles is also strongly affected by the presence
of obstacles. The high-order traveling bands -- reported to emerge in the classical,
homogeneous, Vicsek model [30,31] -- become less pronounced and even disappear
as the level of heterogeneities, i.e. obstacles, is increased. At the point where bands
10 Information spreading in the context of active particles (without alignment) was studied
with self-propelled disks that exchange their internal state upon collision [50]. As for align-
ing active particles, there are also two information mechanisms, with particle-particle infor-
mation exchange being way faster than cluster-cluster information exchange. Interestingly,
fluctuations play a major role even in the absence of cluster-cluster information exchange in
two dimensions where the physics of the problem is always non-trivial [51].
Will be inserted by the editor
15
are no longer observed, the ordering properties change from long-range to quasi-long
range, which suggests that in the limit of an infinite system (keeping constant both,
obstacle and particle density) and for static obstacles, the SPPs cannot maintain
a coherent migratory route along the (infinite) heterogeneous space: i.e. the system
becomes disordered. On the other hand, our finite size analysis revealed that at high
obstacle densities (i.e. ρo ≥ 0.1), the system exhibits two critical points, one at low and
another one at high noise value. Finally, the study of density fluctuations indicated
that the giant-number-fluctuation exponent β moves towards 1/2, which corresponds
to normal density fluctuations, as we approach the noiseless limit, as well as for large
enough obstacle densities.
Finally, it is worth mentioning that few experiments with active particle in het-
erogeneous media have been already performed, so far, with active particles without
alignment: self-propelled janus particles moving on patterned surfaces [46] and speckle
light fields [47,52]. Interestingly, patterned regular environments have been initially
used to rectify the motion of active swimmers such as bacteria in diluted suspensions
(i.e. in a non-interacting context) [53,54,55]. In these systems, volume exclusion ef-
fects and the size of the moving active particles play a central role. Such observations
have triggered a good deal of theoretical work. For instance, in simulations with self-
propelled disks (SPD) it has been shown that SPDs can get locally jammed by the
obstacles [56] 11, while in simulations with SP rods it has been found that V-shaped
obstacles can be used to trap particles [57]. On the other hand, it has been shown
in simulations with circularly moving active particles that a regular configuration of
obstacles can be used to filter the active particles [58,59], while narrow channels can
direct particle motion [60]. It remains to be seen how the large-scale collective prop-
erties reported here are affected by introducing a velocity alignment mechanism in
the above mentioned more realistic models.
References
1. Schaller V, Weber C, Semmrich C, Frey E and Bausch A 2010 Nature 467 73
2. Zhang H, Be'er A, Florin E L and Swinney H 2010 Proc. Natl. Acad. Sci. USA 107
13526
3. Peruani F, Starruss J, Jakovljevic V, Sogaard-Andersen L, Deutsch A and Bar M 2012
Phys. Rev. Lett. 108 098102
4. Buhl J, Sumpter D J T, Couzin I D, Hale J J, Despland E, Miller E R and Simpson S J
2006 Science 312 1402 -- 1406
5. Romanczuk P, Couzin I and Schimansky-Geier L 2009 Phys. Rev. Lett. 102 010602
6. Tunstrøm K, Katz Y, Ioannou C C, Huepe C, Lutz M J and Couzin I D 2013 PLoS
Comput Biol 9 e1002915
7. Holdo R M, Fryxell J M, Sinclair A R E, Dobson A and Holt R D 2011 PLoS ONE 6
e16370
8. Deseigne J, Dauchot O and Chat´e H 2010 Phys. Rev. Lett. 105 098001
9. Kudrolli A, Lumay G, Volfson D and Tsimring L 2006 Phys. Rev. E 74 030904(R)
10. Weber C, Thueroff F and Frey E 2013 arXiv:1301.7701
11. Jiang H R, Yoshinaga N and Sano M 2010 Phys. Rev. Lett. 105 268302
12. Golestanian R 2012 Phys. Rev. Lett. 108 038303
13. Theurkauff I, Cottin-Bizzone C, Palacci J, Ybert C and Bocquet L 2012 Phys. Rev. Lett.
108 268303
14. Palacci J, Sacanna S, Steinberg A, Pine D and Chaikin P 2013 Science 339 936
11 Interestingly, SPD also exhibit an optimal noise that maximizes particle flux [56]. The
explanation for this optimal noise seems to be rooted in the local jamming dynamics, being
intrinsically different from the one related to the optimal noise value reported here.
16
Will be inserted by the editor
15. Golestanian R 2009 Phys. Rev. Lett. 102 188305
16. et al W P 2004 J. Am. Chem. Soc. 126 13424
17. Mano N and Heller A 2005 J. Am. Chem. Soc. 127 11574
18. Ruckner G and Kapral R 2007 Phys. Rev. Lett. 98 150603
19. Howse J, Jones R, Ryan A, Gough T, Vafabakhsh R and Golestanian R 2007 Phys. Rev.
Lett. 99 048102
20. Golestanian R, Liverpool T B and Ajdari A 2005 Phys. Rev. Lett. 94(22) 220801
21. Bricard A, Caussin J B, Desreumaux N, Dauchot O and Bartolo D 2013 Nature 503
95 -- 98 ISSN 0028-0836
22. Peruani F and Morelli L 2007 Phys. Rev. Lett. 99 010602
23. Romanczuk P, Bar M M, Ebeling W, Lindner B and Schimansky-Geier L 2012 Eur.
Phys. J. Special Topics 202 1
24. Dworkin M 1993 Myxobacteria II (Amer Society for Microbiology)
25. Marchetti M C, Joanny J F, Ramaswamy S, Liverpool T B, Prost J, Rao M and Simha
R A 2013 Rev. Mod. Phys. 85 1143 -- 1189
26. Ramaswamy S 2010 Annual Review of Condensed Matter Physics 1 323 -- 345
27. Vicsek T and Zafeiris A 2012 Physics Reports 517 71 -- 140
28. S Ramaswamy R A S and Toner J 2003 Europhys. Lett. 62 196 -- 202
29. Vicsek T, A Czirok E, Jacob E B, Cohen I and Shochet O 1995 Phys. Rev. Lett. 75 1226
30. Gr´egoire G and Chat´e H 2004 Phys. Rev. Lett. 92 025702
31. Caussin J B, Solon A, Peshkov A, Chat´e H, Dauxois T, Tailleur J, Vitelli V and Bartolo
D 2014 Phys. Rev. Lett. 112 148102
32. Toner J and Tu Y 1995 Physical Review Letters 75 4326 -- 4329
33. Toner J and Tu Y 1998 Phys. Rev. E 58 4828
34. Chepizhko O, Altmann E G and Peruani F 2013 Phys. Rev. Lett. 110(23) 238101
35. Peruani F, Deutsch A and Bar M 2008 Eur. Phys. J. Special Topics 157 111
36. Chepizhko O and Peruani F 2013 Phys. Rev. Lett. 111(16) 160604
37. Binder K 1997 Reports on Progress in Physics 60 487
38. Chat´e H, Ginelli F, Gr´egoire G and Raynaud F 2008 Phys. Rev. E 77 046113
39. Kosterlitz J M and Thouless D J 1973 Journal of Physics C: Solid State Physics 6 1181
40. Dey S, Das D and Rajesh R 2012 Phys. Rev. Lett. 108(23) 238001
41. Peruani F, Deutsch A and Bar M 2006 Phys. Rev. E 74 030904(R)
42. Peruani F, Schimansky-Geier L and Bar M 2010 Eur. Phys. J. Special Topics 191 173 --
185
43. Peruani F and Bar M 2013 New J. Phys. 15 065009
44. Ginelli F, Peruani F, Bar M and Chat´e H 2010 Phys. Rev. Lett. 104 184502
45. Peshkov A, Aranson I, Bertin E, Chate H and Ginelli F 2012 Phys. Rev. Lett. 109
268701
46. Volpe G, Buttinoni I, Vogt D, Kummerer H J and Bechinger C 2011 Soft Matter 7
8810 -- 8815
47. Paoluzzi M, Leonardo R D and Angelani L 2014 Journal of Physics: Condensed Matter
26 375101
48. Meschede M and Hallatschek O 2013 New Journal of Physics 15 4
49. Toner J and Tu Y and Ramaswamy S 2005 Ann. Phys. 318 170-244
50. Peruani F and Sibona G 2008 Phys. Rev. Lett. 100 168103
51. Peruani F and Lee CF 2013, Europhys. Lett. 102 58001
52. Volpe G, Volpe G and Gigan S 2014 Scientific Reports 4 3936
53. Galajda P, Keymer J, Chaikin P and Austin R 2007 J. Bacterial. 189 8704
54. Wan M, Reichhardt C O, Nussinov Z and Reichhardt C 2008 Phys. Rev. Lett. 101
018102
55. Tailleur J and Cates M 2009 Europhys. Lett. 86 60002
56. Reichhardt C and Olson Reichhardt C J 2014 Phys. Rev. E 90(1) 012701
57. Kaiser A, Wensink H and Lowen H 2012 Phys. Rev. Lett. 108 268307
58. Mijalkov M and Volpe G 2013 Soft Matter 9 6376 -- 6381
59. Reichhardt C and Reichhardt C J O 2013 Phys. Rev. E 88(4) 042306
60. Radtke P and Schimansky-Geier L 2012 Phys. Rev. E 85 051110
|
1403.6450 | 3 | 1403 | 2016-01-15T15:23:10 | Simulated annealing approach to vascular structure with application to the coronary arteries | [
"physics.bio-ph",
"q-bio.TO"
] | Does the complex processes of angiogenesis during organism development ultimately lead to a near optimal coronary vasculature in the organs of adult mammals? We examine this hypothesis using a powerful and universal method, built on physical and physiological principles, for the determination of globally energetically optimal arterial trees. The method is based on simulated annealing, and can be used to examine arteries in hollow organs with arbitrary tissue geometries. We demonstrate that the approach can generate in-silico vasculatures which closely match porcine anatomical data for the coronary arteries on all length scales, and that the optimised arterial trees improve systematically as computational time increases. The method presented here is general, and could in principle be used to examine the arteries of other organs. Potential applications include improvement of medical imaging analysis and the design of vascular trees for artificial organs. | physics.bio-ph | physics | Simulated annealing approach to vascular structure with
application to the coronary arteries
Jonathan Keelan,1 Emma M. L. Chung,2 and James P. Hague1, ∗
1Department of Physical Sciences, The Open University, Milton Keynes, UK, MK7 6AA
2Department of Cardiovascular Sciences,
University of Leicester, UK, LE1 5WW
(Dated: July 19, 2018)
Abstract
Does the complex processes of angiogenesis during organism development ultimately lead to a
near optimal coronary vasculature in the organs of adult mammals? We examine this hypothesis
using a powerful and universal method, built on physical and physiological principles, for the
determination of globally energetically optimal arterial trees. The method is based on simulated
annealing, and can be used to examine arteries in hollow organs with arbitrary tissue geometries.
We demonstrate that the approach can generate in-silico vasculatures which closely match porcine
anatomical data for the coronary arteries on all length scales, and that the optimised arterial trees
improve systematically as computational time increases. The method presented here is general, and
could in principle be used to examine the arteries of other organs. Potential applications include
improvement of medical imaging analysis and the design of vascular trees for artificial organs.
PACS numbers: 87.10.Rt, 47.63.Cb, 87.19.rm, 87.19.U-
6
1
0
2
n
a
J
5
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
3
v
0
5
4
6
.
3
0
4
1
:
v
i
X
r
a
∗ Author to whom correspondence should be sent, e-mail:[email protected]
1
I.
INTRODUCTION
Arterial trees are vital for the efficient transport of oxygen and nutrients to tissue. Their
anatomy has been studied for many centuries through the dissection of cadavers, inspec-
tion of corrosion casts, medical imaging techniques, and computational models.
It has
been determined that individual arterial bifurcations follow optimality principles that lower
metabolic demand locally [1–6], as demonstrated by the scaling laws followed by arterial
trees [7–11]. More recently, there has been a high level of interest in models that mimic
arterial growth (angiogenesis) using physical and physiological principles to simulate vascu-
lar anatomy. These models are created based on local optimisation principles, where the
anatomy of each branch in the arterial tree is governed by a compromise between max-
imising fluid dynamical efficiency and minimising the quantity of blood required. However,
models of coronary vasculature, based on local optimisation are not able to explain if the
organisation of major arteries is the result of fluid dynamical optimisation across the 'whole
organ' [12–16].
The relationship between the radii of vessels in individual bifurcations is well categorised
by the equation: rγ
p = rγ
d1
+ rγ
d2
where rp is the radius of the parent artery, rd1,2 those
of the daughter arteries, and γ is the bifurcation exponent[4, 17, 18], which has the value
3.0 in Murray's formulation. This relation arises from the combination of the continuity
equation and the diameter-flow rate relation[9]. Several current methods for in-silico growth
of vascular trees into a simulated tissue substrate aim to optimise the local properties of
individual bifurcations [15, 19]. A procedure, known as constrained constructive optimisa-
tion (CCO) starts by inserting a single artery into the tissue. A new vessel with a position
chosen at random is then connected to the original artery and the link point is moved such
that the energy of the arteries is minimised. New arteries are then iteratively added and
optimised until a predetermined number of terminal sites have been added. The overall
result is that CCO and similar methods create trees whose structure is predetermined by
the order in which new arteries are added: if the order is changed, the final tree structure
also changes. Morphologically, CCO reproduces a reasonable distribution of vessel sizes due
to the application of Murray's law, but creates arterial branches that are more symmetrical
than those found in nature (especially for the largest arteries) [20] and significant extensions
are required to generate vessels in hollow organs [21]. The variations in the structure and
2
positions of larger arteries in CCO generated trees are problematic, since organs such as
the heart exhibit only small differences in large artery structure over a population (aside
from rare abnormalities). An alternative method known as global constructive optimisa-
tion (GCO) attempts to overcome the problems of CCO by including a multiscale pruning
update that is global in the sense that it acts simultaneously on a significant subset of the
tree, but otherwise only includes updates allowing local modifications to the topology of the
tree. As such, GCO is limited to sampling a subset of the allowed topologies of the arterial
trees [22], so while it is expected to offer improvements over CCO, it carries no guarantee
of reaching the global minimum. Due to the use of local downhill searches, GCO also has
similar issues with hollow organs. To obtain a universal optimisation technique to compute
in-silico arterial trees for arbitrary tissue structures, a different approach is needed.
Another method for the generation of large scale arterial trees uses extensive morpholog-
ical databases[14, 23]. These trees contain far more vessels than is feasible to generate with
techniques such as CCO, as the topology of the tree is taken from experimental data. How-
ever, since detailed morphological databases do not exist for the vast majority of organs, the
use of these techniques is impossible in the general case. Morphologically generated models
provide trees suitable for large scale fluid dynamical studies and organ phantoms[24]. They
achieve this by reproducing experimental data in a computationally accessible form. As
such they have no predictive powers that can contribute to the understanding of the origins
of arterial tree structure.
A separate class of models exist for use in modeling the growth of tumorous vasculature
and the process of vascular remodeling, which involve the direct simulation of sprouting
angiogenesis[25–28]. These models seek to reproduce the rapid and dynamic process of
tumor vascularisation, or the growth of vasculature in normal tissue as it grows. Using
in silico simulation of sprouting angiogenesis to obtain the vascular structure in a fully
grown organ would be extremely difficult, as full details of the distribution of tissue and
oxygen demands would be needed for all stages of embryonic and childhood development.
As organs such as the heart have very small levels of variation in vascular structure over
the population, the structure itself is likely to be caused by a process different to that of
tumor vascularisation. We suggest that this process is an optimum seeking one, and that as
such an optimisation procedure is required to accurately model it. We examine if, regardless
of the complex processes that guide angiogenesis during growth, the final structure of the
3
coronary vasculature in adult mammals is near optimal.
Development of a method which reaches a morphologically accurate solution based solely
upon optimisation criteria would be useful in vascular research, allowing for the modeling
of realistic vascular trees in organs lacking extensive morphological databases. The inability
of CCO and extensions to find the global energy minimum, and the subsequent lack of
consistent structure (particularly of the larger arteries), is problematic if organ specific
vasculature is required. An approach capable of producing an arterial tree, which minimises
pumping power and blood volume, whilst providing adequate blood flow to critical regions
would be invaluable in this regard. This paper goes beyond previous work by introducing a
far more flexible and universal method for generation of 'whole organ' arterial trees, in any
arbitrarily shaped tissue substrate, that obey both local and global optimisation criteria. To
identify globally optimised arterial trees, we use a powerful computational technique known
as Simulated Annealing (SA) [29]. Although SA is computationally expensive, correctly
applied SA techniques have a key advantage of being mathematically and computationally
proven to converge to a global energy minimum. To achieve this, our SA based approach
has the potential to sample all possible arterial tree configurations, ranging from perfectly
symmetric, intricately bifurcating structures, to asymmetric trees characterised by a single
trunk vessel. This is achieved by allowing: (1) repositioning of bifurcations, and (2) swapping
the parent vessels of bifurcations between different parts of the tree. By introducing these
forms of plasticity to our models, the entire parameter space of the tree can be explored,
allowing the method to identify the best possible arterial configuration for supplying a
particular organ. Full details of this novel method can be found at the end of the article. As
an example application we determine the near optimal configuration of arteries for supplying
the heart and compare our computer generated coronary vasculature with morphological
data from real coronary arteries. Specifically, we determine that the observed anatomy
of the coronary arteries is similar to that expected from near global minimisation of total
energy expenditure, and validate the approach against porcine data, finding a very high
level of agreement with morphological data.
4
II. METHOD
The main purpose of any arterial tree is to maintain adequate blood perfusion with
minimal total metabolic expense. The suitability of an arterial tree for this purpose is
governed by two considerations: (1) since blood is viscous, the power required to pump
blood through the vasculature should be minimised, (2) as energy is required to generate
and maintain blood, the volume of blood required should be minimised. Murray's law
achieves this for individual bifurcations, but the optimal organisation of large numbers of
connected bifurcations is far from obvious. The interplay between these competing concerns
for thousands of arterial segments leads to a complex optimisation problem. Note that in the
following, bifurcations will be referred to as nodes, arteries will be referred to as 'segments
between nodes', and terminal arterioles are referred to as 'end nodes'.
A. Metabolic cost to maintain blood volume
The first component of the approach involves calculating the power needed to maintain
the entire tree, which will be used as a value in the cost function. The power consumption
of the tree can be split into two separate parts: the first is the metabolic cost of maintaining
the blood volume and tissue associated with the tree, and the second is the power required
to pump blood through the tree. The length and radius of each segment (vessel) i of the tree
must be known to calculate the volume. By assuming a fixed bifurcation exponent, the radii
are determined by the topology and only vessel lengths rely on the geometrical arrangement.
To calculate the cost, volume must be multiplied by a constant, mb, corresponding to a
physiologically reasonable metabolic demand of the same quantity of blood and vascular
tissue [30]. Thus the metabolic cost due to the volume of the tree will be given by:
where mb is taken to be 641.3 J s−1 m−3 and Vtree is the volume of the entire tree.
Cv = mbVtree
(1)
B. Power cost to pump blood through vessels
To calculate the power needed to pump blood through the entire tree, we must know the
pressure and volumetric flows inside each segment (vessel) of the tree, which can be found
5
by first assuming that Poiseuille's law, ∆p = QR, is followed inside the segments, where ∆p
is the pressure drop over the vessel, and Q is the flow. The assumption that flow is laminar
inside the vessels is justified provided that the typical length of a vessel is much larger than
the radius, and that pulsatile flow effects are negligible. Vessels within the simulated trees
have a typical length radius ratio of 10, and while in the largest arteries of the tree pulsatile
effects may still be present, these rapidly decay so that the vast majority lie within a non-
pulsatile regime. We assume both Murray's law and that terminal node flows are constant
to simplify calculation of the relevant fluid dynamical quantities: the only quantity which
relies on the structure of the tree is the pressure. In a sense, the segments can be considered
as connected set of resistors, with the resistance given by:
R =
8µL
πr4 ,
(2)
where r is the radius of the vessel, L its length and µ = 3.6× 10−3Pa s the viscosity of blood.
The pressures (and hence flows) for every node in the tree can then be found recursively.
Wi, the power consumed by each segment i is then calculated using:
Wi = Q2
i Ri,
(3)
Summing over all segments in the tree, the total power required to maintain the proper flow
through the tree is:
Cw =
Ntot(cid:88)
Wi.
(4)
C. Ensuring tissue supply
i
The primary purpose of the vascular tree is to supply blood, thus it is important that
terminal nodes are correctly dispersed inside the tissue. Initially, terminal nodes are ran-
domly distributed inside the tissue, with each node having associated with it a sphere of
influence for blood supply. The radius of this sphere is calculated using physiological values
for the blood demand of the tissue. The density of myocardium is ρ = 1.06 × 103kg m−3
[31], and the flow demand is 1.13 ml min−1g−1 [32] leading to a flow demand per m3 of heart
tissue of qrequired = 2 × 10−2s−1. The total flow into the heart is Q0 = 4.16 × 10−6m3s−1
[33], which can be converted to total flow per node as QN = Q0/N , where N is the total
number of arterioles (end nodes). The radius of the supply sphere is then calculated via
6
FIG. 1: (a) The distance map for a spherical surface. Voxels outside the surface have value
0, with those inside the surface contributing a value relating to their distance from the
surface. (b) Each arteriole supplies a spherical region shown by the lightly shaded squares.
Where there is significant overlap between two spheres, there is a penalty. Unsupplied
voxels also incur a penalty in the cost function.
4πR3
supply/3 = QN /qrequired. The sphere can be thought of as a microcirculatory 'black box'
[15], where the exact fluid dynamical details of the blood flow have been ignored. Spheres
of blood supply associated with end nodes are stored in a voxel map (a voxel is a 3D gen-
eralisation of a pixel) of the tissue, where each terminal node adds exactly one to each
voxel inside its sphere of supply (Fig 1b). While blood demand is not constant within the
myocardium at any single instance of time, the majority of fluctuations are high frequency
oscillations which are assumed to be averaged out in the present model[34]. The terminal
nodes are then allowed to move inside the tissue, where after each move a new voxel supply
map is calculated, and the overlap (each voxel supplied by more than 1 sphere , or the dark
red voxels in Fig 1b ) is used as a value in the cost function of the simulated annealing
algorithm. In addition, all voxels not being supplied are given a cost, so that the overall
penalty associated with having both unsupplied and oversupplied voxels is chosen to be:
(cid:88)
voxels
Cs =
s; s =
10 if b = 0
(b − 1)2 otherwise
(5)
where b is the value of the supply at the voxel and the sum is performed over all the voxels
comprising the tissue. In practice, this cost is set to be much larger than all other costs, since
any unsupplied tissue would die. Therefore, the terminal nodes spread evenly through the
tissue early in the optimisation. Other functions may be used, provided that the minimum
in the function for each voxel occurs at b = 1. The value Cs then defines the fitness of the
tree to supply blood, and the penalty for over supplying voxels forms a sort of self avoidance
7
algorithm, where terminal nodes are encouraged to pack the tissue as densely as possible
without overlapping. A benefit of this method is that it allows easy integration of medical
imaging into the model, as well as providing an easy method for differentiating tissue with
different blood supply demands.
D. Exclusion of large vessels from tissue
In order to create a realistic vascular tree, it must be possible to exclude some segments
from penetrating the tissue. For instance, in the case of the heart, it would be unlikely to
find a very large artery within the myocardium, and vessels may not penetrate the ventricles;
rather, the larger arteries and arterioles lie on the surface of the heart, with only the smaller
arterioles and capillaries being found inside the tissue. To mimic this structure, the approach
makes use of a cut off radius Rcutoff, whereby segments with radius larger than Rcutoff may
not penetrate the tissue. In the calculations performed in this article, Rcutoff = 0.01 mm.
To determine which segments with radius greater than Rcutoff have penetrated the tissue we
first take a distance transform of the tissue surface for each tissue voxel (Fig 1a. ) This
provides a second voxel map of the tissue, distinct from the blood supply map, giving a
measure of the distance of a point from the surface when it is inside the tissue (outside of
the surface, the value is zero). For each segment satisfying the radius criteria, a list of voxels
that its centre-line penetrates is generated [35], along with a value for the length element of
the segment present inside that voxel. A cost is then calculated based upon the value of the
distance transform at each of the voxels according to,
Co = πr2(Dijk
Lijk)6,
(6)
where i, j and k are the cartesian voxel coordinates taken from the centre-line of the segment.
Dijk is the value of the distance transform at that voxel coordinate. Lijk is the length of
the segment spent inside the voxel. The sum is performed over all the voxels contained in
the list calculated from the centre-line. This cost can then be used in the SA algorithm as
a penalty that favours moving large segments out of the tissue.
8
E. Pressure constraints
In physiologically realistic trees, capillary networks should receive a constant pressure
Pterm to function correctly. A new cost can be devised to ensure this. A suitable candidate
is,
Cp =
Nterm(cid:88)
(Pi − Pterm)2,
(7)
i
where the sum is performed over all terminal nodes, and Pi is the actual terminal node
pressure.
In practice, for trees which can be optimised on feasible time scales (i.e of a
few thousand nodes), the pressure drop from root to end node is less than 1% of the total
pressure drop of a real arterial tree, with most of the pressure drop occurring over smaller
arterioles than those considered here, so it is unnecessary to perform this calculation. When
it becomes possible to grow larger trees, the pressure at the capillaries will need to be taken
into consideration. This will add a significant computational cost.
F. Total cost function
We have now determined a form for all the relevant costs associated with an arbitrary
tree configuration supplying arbitrary tissue shapes. We can therefore define a total cost
which gives a numeric measure of the fitness of a given tree,
CT = Aw,v(Cw + Cv) + AoCo + ApCp + AsCs
(8)
where Ai indicates a weighting value which scales each relevant cost. There is no way to
analytically determine what weights to use, and the selection of appropriate weights must
found experimentally, however a few basic principles such as the having a very high weight
for the blood supply cost and a low weight for the end node pressure cost can guide the
process. In principle, As should be infinite, since tissue without supply dies. In this work,
we use Aw,v = 1× 104, Ap = 0, As = 1× 1030 and Ao = 100. In this way, As and Ao force the
exclusion of vessels and uniform supply of tissue to act like constraints. While the exclusion
cost should technically be infinite, as no arteries are found in the ventricles of living humans,
it is advantageous to give it a large but finite value. This allows the optimisation procedure
to identify gradients, giving it extra information and speeding up convergence. This size of
the constant for Ao may seem small in this regard, however its value Co is already raised to
9
FIG. 2: Tree modification updates
the power 6 in Eq 6. Since any scaling of the cost function does not effect the location of
minima, we can absorb one of the weightings by scaling everything else. This allows a new
cost function CT = CT
Aw,v
to be defined.
G. Simulated annealing
To select the fittest, most optimised trees, we use a powerful technique for optimisation
problems known as simulated annealing (SA) [36, 37]. The primary difference between SA
and a conventional downhill search is that SA also spends some time exploring solutions
with higher cost function, and in this way can climb out of shallow valleys in the fitness
function to explore other deeper regions.
The total cost CT will play the role of energy in the simulated annealing algorithm, so
T , resulting in a tree of cost C f
that the probability of accepting a change to a tree of cost C i
T
exp(cid:0)− ∆CT
T
(cid:1) if ∆CT > 0
1 otherwise
is given as
P f
i =
(9)
Where P f
i
is the probability of going from state i to state f , ∆CT = C f
T is the change
in the cost function associated with going from state i to f , and T is the simulated annealing
T − C i
temperature parameter (not be confused with ambient temperature). The small probability
to accept a higher cost tree during update allows the tree to climb out of local valleys in the
cost function. The algorithm proceeds by making changes to the tree structure, calculating
the change in cost function, and then either accepting or rejecting the change by comparing
10
P f
i to a random number between 0 and 1. T starts large and is reduced slowly. If T has been
reduced sufficiently slowly, then the global minimum of the cost function is guaranteed to be
reached. In practice, the problem space is too large to achieve this in reasonable time, and
slightly different trees with very similar cost are found if the algorithm is run with several
random number seeds. The most important consideration is the lowest achieved cost. As
such, if the structure of trees generated varies between different runs, we always display data
from the run with the lowest cost function. As computational power increases, longer runs
will be achievable leading to progressively better optimisations. The highest T used here is
1× 1010, dropping during the algorithm to 10−5. Typically a tree containing 1000 nodes will
need 109 updates, with a doubling of nodes taking roughly quadruple the number of updates
(up to around 6000 Nodes with 1 month of CPU time). The large value of As means that
the supply of tissue is determined by downhill search, while all other costs are minimised by
simulated annealing.
H. Exploring the tree structure: Translations and node swaps
The SA algorithm must have access to set of updates which allow it to alter the configu-
ration of the tree. It is necessary to find changes that can be made to the topological and
geometrical structure of the tree such that all possible solutions, between perfectly symmet-
ric structures and a single trunk vessel can be explored (i.e. the algorithm is ergodic). This
is achieved by allowing: (1) repositioning of bifurcations, which is achieved by translating a
node in space (Fig. 2a) and (2) swapping the parent vessels of bifurcations between different
parts of the tree (Fig. 2b). For all nodes but the root node, this move is valid, and per-
formed consecutively it allows all possible tree topologies to be explored. If one of the two
nodes is a direct parent of the other (i.e while traversing up the tree from one of the chosen
nodes, the other node is encountered) the move is rejected to avoid forming a closed loop.
With these two updates, the entire parameter space of the tree can be explored, allowing
the algorithm the opportunity to reach a globally optimal solution.
11
FIG. 3: Schematic of the Strahler ordering process.
I. Strahler Order
The Strahler (or stream) ordering method was first introduced to classify river systems,
but can be applied to any bifurcating system. In standard Strahler ordering, nodes at the
end of a tree (in this case the arterioles) are assigned a number 1. At a bifurcation, if
two vessels (segments) of the same order meet, then the order of the parent vessel is 1
higher. However, if two vessels of different orders meet, the artery supplying these vessels
has the largest order of the two. For example, if two arteries of order 1 meet, then the vessel
supplying these arteries has order 2. If an artery of order 3 meets an artery of order 2, then
the vessel supplying these arteries has order 3 (an example is shown in Fig. 3). Therefore,
within this scheme, vessels with the lowest order are arterioles. The major vessels have the
largest order. The Strahler order used here is then diameter adjusted following the approach
in Ref. [38].
Within the Strahler ordering scheme it is possible to identify continuous sections of vessels
with the same order number. These are refereed to as elements, so a single arterial element
may pass through multiple bifurcations. Throughout this article it is the properties of
elements which will be calculated for direct comparison with Ref. [9]. We note that due to
the early termination of the simulated trees, calculated order numbers are modified so that
the root nodes have an order number equivalent to that of the largest arteries of real coronary
arterial trees. For example, in the work of Kassab, the largest diameter defined Strahler
order number is 11, corresponding to the input artery. For a computer generated tree of
only 6000 nodes spanning order numbers 1-6, 5 must be added to each order number so that
the orders of the root nodes (largest vessels) match and a direct comparison can be made.
This is consistent with assuming that the smallest vessels in the computer generated tree
correspond to vessels of order 6. Which is due to the absence of smaller vessels downstream
12
FIG. 4: Images showing arterial trees grown with the approach detailed here. The number
of terminal arterioles is increased from 500 to 6000 (the total number of arterial segments
is roughly twice this). There is consistency in the positioning of the larger arteries between
the numerical method and the typical arrangement of the major arteries, suggesting that
the coronary arteries may be the result of a biological process seeking the global minimum
in metabolic demand.
of the smallest arteries in the in-silico model.
III. RESULTS
In this paper globally optimised vessels are grown using an SA based approach to supply
a myocardial substrate, and validated through comparison with morphological data from
the porcine arterial tree. We choose to examine the heart vasculature, since the structure
of the large coronary arteries has been found to be similar between individuals [40] and
the full arterial tree has been well characterised in porcine models [9]. For modelling the
coronary arteries we used the following parameters: (1) A tissue substrate representing an
ellipsoidal human heart muscle of mass 218g, constructed based on physiological parameters
[41] . The right ventricle was assumed to take the form of a super ellipsoid of exponent 2.5
and the left ventricle was represented by a simple ellipsoid. Truncation of the ellipsoidal
substrate was chosen so that the mass of the tissue corresponded to a reasonable physiological
13
(a)
(b)
(c)
(d)
FIG. 5: (a) Vessel diameter as a function of order in a tree with 6000 arterioles. Excellent
agreement is found for vessels on all length scales. (b) Vessel length as a function of order
number. Agreement is excellent for the major vessels (large order). The large variation
seen for arterioles (lower order) is a result of early termination. Also shown are the
morphological data reproduced from Table 2 of Ref. [5] for easy comparison. (c) and (d)
are as (a) and (b), but for smaller trees to highlight the trend towards the morphological
data as tree size increases. (Error bars show standard errors, both axes are logarithmic.)
value given morphological data for ventricle thickness. (2) Blood flow through each of the
terminal segments of the tree was assumed to be constant, with each arteriole supplying
an equal volume of tissue and homogeneous perfusion throughout the tissue parenchyma
[42]. These assumptions greatly simplify fluid dynamical calculations for estimating the
total power needed to pump blood through the tree. (3) The metabolic cost of maintaining
a given volume of blood was assumed to be 641.3J s−1 per metre cubed of blood [30] . For
14
4681012101102103104Order NumberDiameter µ m Morph DataMorph FitGenerated24681012101102103104105Order NumberLength µ m Morph DataMorph FitGenerated3456789101112101102103104Order NumberDiameter µ m Morph Fit500 Arterioles1000 Arterioles2000 Arterioles24681012101102103104105106Order NumberLength µ m Morph Fit500 Arterioles1000 Arterioles2000 Arterioles(a)
(b)
FIG. 6: The ratio of daughter vessel diameters (Ds and Dl are the diameters of the
smallest and largest daughter vessels respectively) to diameters of parent segments, Dp as a
function of order number, showing how the tree tends towards more symmetric branching
at lower orders. Agreement with morphological data reproduced from tables in the online
supplement of Ref. [39] is good, if the early termination of the generated trees is taken into
account, with the trend towards the morphological data as the tree size increases. Both
graphs demonstrate that there are large trunks at high orders with the largest daughter
vessel (panel (b)) of similar size to the parent vessel and another side artery which is much
smaller (panel (a)). At smaller orders, the ratio becomes similar showing that the
branchings of the smaller arteries are near symmetric. Realistic branching asymmetries are
a clear advantage over other methods of generating arterial trees in-silico.
convenience, each arteriole supplies a sphere of tissue with a size calculated by assuming a
mean blood flow per unit mass for cardiac muscle of 0.8 ml min−1 g−1[43]. The value taken
from the literature was chosen such that it lay within the given error, but also conformed
reasonably with both the ellipsoidal heart model, input flow and radii.
(4) The larger
arteries with diameters greater than 0.01mm were constrained to avoid penetration of the
outer layer of heart tissue. This simplification differs slightly from real coronary vasculature,
where progressive intrusion of arteries into the myocardium can be observed [44]. However,
as the major arteries modelled by our method are far larger than the intra-myocardial vessels,
a sharp cut-off is thought to provide a reasonable approximation. (5) The starting positions
of the two root arteries were fixed with a total input flow of 4.16−6 m3 s−1[45]. Relative
radii of the two inputs to the tree were constrained via r2.1
2 = [2.1mm]2.1 however,
1 + r2.1
15
24681000.20.40.60.81Order NumberDs / Dp Morph Data2000 Arterioles6000 Arterioles24681000.20.40.60.81Order NumberDl / Dp Morph Data2000 Arterioles6000 ArteriolesFIG. 7: Example trees generated with different values of mb, which changes the relative
weight of the pumping power to cost of maintaining blood in the optimisation. For small
mb (corresponding to small hearts), vessels in the trees wind around - this is because there
is little penalty to make a single wide vessel that curves to supply blood, rather than
bifurcating. For large mb (corresponding to large hearts) the vessels travel as straight as
possible.
the relative sizes of root arteries and division of perfusion territories are determined by the
method alone. (6) The branching exponent varies throughout the coronary arterial tree,
but for the larger arteries its value remains in the range 1.8 to 2.3. A variable branching
exponent would greatly increase the computational cost of the approach, so a compromise
value of 2.1 was chosen for the entire tree [39].
Coronary arterial trees containing increasing total numbers of vessels grown using the SA
based method are presented in Fig. 4. In real human coronary trees, there are 3 identifiable
main coronary arteries (see e.g. the schematic from Ref.
[46]): Left Anterior Descending
(LAD), Right Cardiac Artery (RCA) and Left Circumflex Artery (LCX). The positions and
relative dimensions of these are similar in most humans, with major variations observed in
less than 1% of healthy individuals [47]. Trees grown using SA (Fig. 4) adhere well to this
structure. There is a consistency in the placement of the larger arteries, although the RCA
appears slightly lower, and the right marginal artery appears slightly shorter, in our models.
Overall, visual inspection of the arterial structure appears extremely promising.
To provide a quantitative comparison of our trees with anatomical data, the topological
characteristics of the computer generated coronary artery trees were extracted and compared
to morphological data characterising the pig coronary arteries published by Kassab et al.
[9] Kassab and colleagues used a combination of corrosion casting and optical sectioning
to obtain detailed morphometric data, tabulated using the Strahler (or stream) ordering
16
scheme to denote elements of the tree of varying scale. Within this scheme, the lowest
Strahler order numbers correspond to the smallest arterioles and the largest numbers refer
to major vessels (for details on Strahler ordering see method). To directly compare arterial
diameters, lengths, and branching properties, of our computer-generated arterial tree with
real data from pig coronary arteries, averages were obtained over all elements of the same
diameter defined Strahler order.
The mean vessel diameters are shown as a function of order number, for a tree compris-
ing of 6000 arterioles (12000 vessel segments) in Fig. 5(a). Excellent agreement is found
between the trees generated in-silico and the morphological data. Only slight deviations
from the morphological data can be seen for the smallest vessels (lowest order arteries) in
the generated tree. This is likely to be due to the combination of integer order numbers and
the condition that terminal sites are of constant radius. The result of this constraint is that
the terminal radii will only match the anatomical data for a correct choice of the number of
arterioles. Fig 5(c) shows the effects on diameter of increasing the number of arterioles from
500 to 2000. Agreement is generally good, regardless of the number of terminal arteries,
and there is a clear trend towards matching the experimental data as simulated tree size
increases. Fig 5(b) compares average vessel length in the model and porcine morphologi-
cal data as a function of order number. For the largest arteries (high order numbers) the
agreement is excellent. Although the lengths of the smaller arteries (Strahler orders < 7) in
the computer generated tree tended to be overestimated, this can be easily explained by the
fact that the smallest vessels are required to bridge a gap that would normally be filled by
inclusion of lower order vessels in a larger simulation. As the number of generated vessels is
increased, the agreement with morphological data improves (Fig. 5(d)).
Previously, the best methods available for the computer generation of arterial trees strug-
gled to recreate realistic branching asymmetry. Fig 6 shows the ratio of daughter to mother
vessel radii for the largest and smallest daughter vessels as a function of order number. This
provides a measure of the branching asymmetry of the tree, where small ratios indicate that
branching is symmetric, while ratios approaching 1 suggest a large trunk vessel with small
branches. For Strahler orders corresponding to microvascular arterioles, both the computer
generated and true morphology approach 0.7, which is consistent with perfectly symmetric
branching where both daughter vessels are of similar size. Agreement with the morpho-
logical data from Ref.
[39] improves as the size of the computer generated tree increases.
17
This is not the result of any special input parameters or initial conditions. The trees are
topologically and spatially randomised before SA optimisation begins, and are allowed to
explore the entire parameter space during optimisation. The observed asymmetry is purely
the result of a balance between pumping power and metabolic maintenance cost, and is a
major improvement in predicting the trunk-like structure of major vessels.
Our final figures show the effect of altering the metabolic energy cost of blood per unit
volume mb. The largest morphological change is found in the lengths of the larger arteries
(Fig 8). As mb increases, bifurcation symmetry is also increased in the larger arteries and
as a result there is an increase in the number of Strahler orders present in the tree (Fig 8).
The explanation for these scaling behaviours is evident when considering the limiting cases.
For mb = 0 the power involved in pumping the blood dominates the optimisation, which
leads to a large, 'snaking' artery with small side branches that supply the tissue. This large
artery would cover the entire surface of the heart, and the configuration is equivalent to a
completely asymmetric binary tree. For a large mb value (or small power cost) there is a
huge penalty associated with larger arteries, and so their lengths are contracted. In order
to accommodate the reduction in length, the larger arteries must bifurcate more frequently
and symmetrically. Additionally the high volume cost causes the trunk artery to minimise
its total length, resulting in a much straighter path across the tissue. Less extreme examples
of this behaviour can been seen in Fig 7, with meandering arteries for small mb and straight
arteries for large mb.
The change in mb can also be interpreted as a change in length scale as follows: Once
the large vessels have been excluded from the tissue and all tissue is supplied, the remaining
cost function that is optimised has the form,
now, make the transformations, r → r(cid:48) = Ar, l → l(cid:48) = Al. Then the cost function becomes,
C = mbπr2l +
8µlQ2
πr4
C = A3mbπr2l +
8µAlQ2
A4πr4
since the optimum in the cost function is the same independent of a multiplicative factor
that acts on all terms, then we can absorb a factor of 1/A3 into the cost function to obtain:
C(cid:48) = A6mbπr2l +
8µlQ2
πr4
18
identifying a new m(cid:48)
b = A6mb the cost function now has the same form. Since changing mb
is equivalent to changing the length scale, these results suggest that there are likely to be
structural differences between species of different sizes, as the power required to pump blood
becomes relatively more important than the metabolic demand to maintain blood volume
in small vessels. In the absence of morphological data, visual comparison of the coronary
arteries tentatively indicates that vessels meander around in smaller species [48] and that
vessels are straighter in larger species [49].
IV. DISCUSSION
We have developed a powerful and universal method for growing arterial trees in-silico,
which is capable of identifying the near globally optimal configuration of arteries for arbi-
trarily shaped tissues with heterogenous blood supply demands. As input, the method only
needs information about the tissue structure and the entry point positions of the largest
arteries. From this information, the approach generates morphologically and structurally
accurate coronary arterial trees at almost every length scale. This is a significant improve-
ment on previous optimisation methods, which failed to reproduce the consistent structure
found in the coronary arteries. We have shown that the method improves with the number
of vessels modeled, so that, as computing power increases, there is a systematic improvement
in the accuracy of the generated trees. To our knowledge, no other method can generate
realistic arterial trees that closely match morphological data by taking only the shape of
the tissue as input, and claim systematic improvement in the generated trees with increased
computational power.
We expect that our method could have several useful applications. Our first application
is to use cardiac and partial cerebral vasculature structures (e.g. MCA territory) computed
using this method as input to models of embolic stroke and other infarctions, since these
models require detailed vasculature structure over a range of length scales which are not
available to imaging techniques [50–52]. This will require further validation of the algorithm
against cerebral arterial data. Such vasculature would be downstream of the Circle of Willis,
a source of major anatomical variation and likely not a structure reproducible by the current
algorithm. Computational models of stroke combined with doppler ultrasound have potential
to provide further information regarding embolic burden during major operations [53].
19
There are several other potential applications. Models of arterial trees generated by our
method may help to improve the interpretation of medical images though advanced image
segmentation techniques. Vessels identified through automatic segmentation techniques can
be connected via algorithms such as the one presented[54]. In addition, segmentation can
be limited to the location of a reduced set of bifurcation points, and the algorithm used to
fill in any missing vessels which connected them [55].
We also speculate that the algorithm could be of use in designing the structures of vascu-
lature for artificial tissues. Once the very difficult and intricate process of making networks
of vessels in artifical tissue has been achieved[56–58], the opportunity to optimise or inform
their design will be available. A common problem during the growth of artificial tissue is
that regions of cells can die due to lack of nutrients and oxygen. For instance an artificial
skin graft may be optimised for increased healing, by having its vasculature designed such
that there is higher overall perfusion with minimal loss of useful tissue. Even more specu-
latively, artificially grown organs may have a vasculature designed to minimise their impact
on the cardiovascular system.
V. ACKNOWLEDGMENTS
We thank Chloe Long, Martin Bootmann and Uwe Grimm for useful discussions.
VI. DATA ACCESSIBILITY
Data presented in this articles can be found in a compressed zip folder as Electronic
Supplementary material. The folders contain MATLAB scripts capable of plotting the data.
VII. COMPETING INTERESTS
We have no competing interests.
VIII. AUTHORS' CONTRIBUTIONS
JK developed the algorithm, acquired, analysed and interpreted data, and contributed to
drafting the article. EC co-conceived the study, co-supervised the project and contributed
20
to drafting the article. JPH conceived the study, developed initial versions of the algo-
rithm, contributed to the analysis and interpretation of data, supervised the project, and
contributed to drafting of the article. All authors gave final approval for publication.
IX. FUNDING
JK acknowledges EPSRC grant EP/P505046/1. EMLC acknowledges support from a
British Heart Foundation Intermediate Basic Science Research Fellowship (FS/10/46/28350).
Appendix A: Convergence and Consistency
The primary purpose of the algorithm is to produce arterial tree configurations which
conform to those found in living organisms. As the algorithm itself relies only upon opti-
misation principles, the close agreement with experimental results implies an evolutionary
pressure towards a structure with minimal power consumption. This is itself a far from new
concept, however in this paper we have shown that energetic constraints lead not only to a
morphometrically realistic tree, but also to the production of major arteries which closely
follow the paths of major arteries in living systems.
Whilst it is clear that the algorithm produces both morphometrically and geometrically
realistic structures, what is not clear is how close the optimisation procedure gets to the
global energy minimum, or indeed whether there is a non-degenerate energy minimum at all.
For any given topological configuration there is a single, non-degenerate energy minimum
which it is possible to approach using Newton-Rhapson (provided the solution space is
convex, which would not be true for more complex structures). In contrast, the topological
space for any even modestly sized tree is huge, highly degenerate and not easily searchable.
Even if one excludes degenerate topological structures, which in the case of the swap node
procedure outlined earlier would imply never swapping two nodes with the same number
of distal terminal sites, the number of possible configurations is still massive. It is entirely
possible that two distinct topological configurations share a degenerate energy level, and
proving that this is not the case appears difficult.
While it may be that the global energy minimum is highly degenerate, the algorithm
itself can still be characterised in terms of reliability and convergence. For reliability, we can
21
perform visual inspections on trees and assess their similarity. It must be noted here that
the geometry in which the tree is grown will have a large effect on the consistency of the
results. For instance, in the case of a circular section of tissue with an input in the centre,
there is a high degree of rotational symmetry. In the case of convergence, we can produce
many trees and plot the frequency distribution of their resultant energies, or as in this case
the average and variance of the cost as a function of SA steps.
The convergence and consistency test trees were generated on a 2D plane with the input
placed in one corner. The trees consisted of 127 nodes total (64 end nodes) and had a
bifurcation exponent of 3.0. The 2D tissue plane was sized at 10cm by 10cm and the root
radius at 2.4mm. Each tree was optimised for a given number of simulated annealing steps,
with the minimum energy of the SA run being recorded. The average energy reached for
a given number of SA steps was then calculated (Fig. 9). The results show a clear trend
towards lower average energy and standard deviation as the number of SA steps increases.
The high variance at the lower numbers of SA steps are typical of a system which has been
quenched, i.e high temperature disorder has been locked into the system, which had not had
sufficient time to reach equilibrium.
For the consistency test we have a produced Fig. 10, which show trees generated for three
different numbers of steps. As would be expected, at low numbers of SA steps the trees are
very dissimilar, however as the number of steps is increases the similarity between the overall
trees increases dramatically, with a main diagonal artery dominating the structure.
[1] M. Zamir. The Physics of Pulsatile flow. Springer, 2000.
[2] MDS Frame and IH Sarelius. Microvascular research, 50:301–310, 1995.
[3] C D Murray. Proceedings of the National Academy of Sciences of the United States of America,
12(3):207–14, March 1926.
[4] S Rossitti and J Lofgren. Stroke: A Journal of Cerebral Circulation, 24(3):371–377, 1993.
[5] G S Kassab, C A Rider, N J Tang, and Y C Fung. American Journal of Physiology,
265(1):H350–65, 1993.
[6] F Cassot, F Lauwers, S Lorthois, P Puwanarajah, V Cances-Lauwers, and H Duvernoy. Brain
research, 1313:62–78, February 2010.
22
[7] Y Huo and G S Kassab. Biomedical Engineering, 9:190–200, 2012.
[8] Y Huo and G S Kassab. Biophysical journal, 96(2):347–53, January 2009.
[9] G S Kassab. American journal of physiology. Heart and circulatory physiology, 290(2):H894–
903, February 2006.
[10] H U Bengtsson and P Ed´en. Journal of theoretical biology, 221(3):437–43, April 2003.
[11] G B West, J H Brown, and B J Enquist. Science, 276(5309):122–126, 1997.
[12] R Karch, F Neumann, M Neumann, and W Schreiner. Computers in biology and medicine,
29(1):19–38, January 1999.
[13] R Karch, F Neumann, M Neumann, and W Schreiner. Annals of Biomedical Engineering,
28(5):495–511, May 2000.
[14] B Kaimovitz, Y Lanir, and G S Kassab. Annals of biomedical engineering, 33(11):1517–35,
November 2005.
[15] W Schreiner and P F Buxbaum. IEEE transactions on bio-medical engineering, 40(5):482–91,
May 1993.
[16] B Kaimovitz, Y Lanir, G S Kassab, S Nees, G Juchem, N Eberhorn, M Thallmair, S Forch,
M Knott, A Senftl, T Fischlein, B Reichart, D R Weiss, and H G M Van Beek. Am J. Phys.
Heart Circ. Phys., 299(July 2010):1064–1067, 2010.
[17] H N Mayrovitz and J Roy. Microvascular blood flow: evidence indicating a cubic dependence
on arteriolar diameter. The American journal of physiology, 245(6):H1031–H1038, 1983.
[18] Horsfield K. and Woldenberg MJ. Diameter and cross-sectional areas of branches in the human
pulmonary arterial tree. Anat Rec., 223(3):245–251, 1989.
[19] R Karch, F Neumann, M Neumann, and W Schreiner. Annals of biomedical engineering,
28(5):495–511, May 2000.
[20] W Schreiner, F Neumann, M Neumann, R Karch, A End, and S M Roedler. The Journal of
general physiology, 109(2):129–140, 1997.
[21] W Schreiner, R Karch, M Neumann, F Neumann, P Szawlowski, and S Roedler. Medical
Engineering & Physics, 28(5):416–429, 2006.
[22] HorstK. Hahn, Manfred Georg, and Heinz-Otto Peitgen. In Gabriele A. Losa, Danilo Merlini,
Theo F. Nonnenmacher, and Ewald R. Weibel, editors, Fractals in Biology and Medicine,
Mathematics and Biosciences in Interaction, pages 55–66. Birkhauser Basel, 2005.
[23] G S Kassab, E Pallencaoe, A Schatz, and Y C Fung. American Journal of Physiology, 273(6
23
Pt 2):H2832–H2842, 1997.
[24] George S K Fung, W Paul Segars, Grant T Gullberg, and Benjamin M W Tsui. Development
of a model of the coronary arterial tree for the 4d xcat phantom. Physics in Medicine and
Biology, 56(17):5651, 2011.
[25] Holger Perfahl, Helen M. Byrne, Tingan Chen, Veronica Estrella, Toms Alarcn, Alexei Lapin,
Robert A. Gatenby, Robert J. Gillies, Mark C. Lloyd, Philip K. Maini, Matthias Reuss, and
Markus R. Owen. Multiscale modelling of vascular tumour growth in 3d: The roles of domain
size and boundary conditions. PLoS ONE, 6(4):e14790, 04 2011.
[26] A.R.A. Anderson and M.A.J. Chaplain. Continuous and discrete mathematical models of
tumor-induced angiogenesis. Bulletin of Mathematical Biology, 60(5):857–899, 1998.
[27] Steven R. McDougall, Alexander R.A. Anderson, and Mark A.J. Chaplain. Mathematical
modelling of dynamic adaptive tumour-induced angiogenesis: Clinical implications and ther-
apeutic targeting strategies. Journal of Theoretical Biology, 241(3):564 – 589, 2006.
[28] Anusuya Das, Douglas Lauffenburger, Harry Asada, and Roger D. Kamm. A hybrid contin-
uum–discrete modelling approach to predict and control angiogenesis: analysis of combina-
torial growth factor and matrix effects on vessel-sprouting morphology. Philosophical Trans-
actions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences,
368(1921):2937–2960, 2010.
[29] S. Kirkpatrick, C. D. Gelatt, and M.P. Vecchi. Science, 220:671–680, 1983.
[30] Y Liu and G S Kassab. American journal of physiology Heart and circulatory physiology,
292(3):H1336–H1339, 2007.
[31] Kalyan C Vinnakota and James B Bassingthwaighte. American journal of physiology. Heart
and circulatory physiology, 286(5):H1742–H1749, 2004.
[32] N G Uren, J A Melin, B De Bruyne, W Wijns, T Baudhuin, and P G Camici. The New
England journal of medicine, 330(25):1782–1788, 1994.
[33] John E Hall. Guyton and Hall Textbook of Medical Physiology. Elsevier, 2010.
[34] R.B King and J.B. Bassingthwaighte. Temporal fluctuations in regional myocardial flows.
Pflugers Arch., 413:336–342, 1989.
[35] J Amanatides and A Woo. Delta, i(3):3–10, 1987.
[36] D Henderson, S H Jacobson, and A W Johnson. In Handbook of metaheuristics, chapter 10,
pages 287–319. Kluwer, 2003.
24
[37] S Kirkpatrick, C D Gelatt, and M P Vecchi. Science (New York, N.Y.), 220(4598):671–80,
May 1983.
[38] Z L Jiang, G S Kassab, and Y C Fung. Journal of Applied Physiology, 76(2):882–892, 1994.
[39] B Kaimovitz, Y Huo, Y Lanir, and G S Kassab. American journal of physiology Heart and
circulatory physiology, 294(2):H714–H723, 2008.
[40] R Glenny, S Bernard, B Neradilek, and N Polissar. Proceedings of the National Academy of
Sciences of the United States of America, 104(16):6858–6863, 2007.
[41] J. J. J. M. Van Den Broek and M. H. L. M. Van Den Broek. J. Biomechanics, 13:493–503,
1980.
[42] A R Pries and T W Secomb. Cardiovascular Research, 81(2):328–335, 2009.
[43] F.J. Klocke, I.L. Bunnell, D.G Greene, S.M. Wittenberg, and J.P. Visco. Circulation,
50(3):547–549, 1974.
[44] S Sunni, S P Bishop, S P Kent, and J C Geer. Archives of pathology laboratory medicine,
110(5):375–381, 1986.
[45] K Johnson, P Sharma, and J Oshinski. Journal of Biomechanics, 41(3):595–602, 2008.
[46] OpenStax College. Anatomy and Physiology [ISBN 978-1-938168-13-0, Available at Connex-
ions Web site, http://cnx.org/content/col11496/1.6/]. June 2013.
[47] A. M. Gharib, V B Ho, D R Rosing, D A Herzka, M Stuber, A E Arai, and R I Pettigrew.
Radiology, 247(1):220–227, April 2008.
[48] A Yoldas, E Ozmen, and V Ozdemir. Journal of the South African Veterinary Association,
81(4):247–252, 2010.
[49] O. Ozgel, A. Haligur, N. Dursun, and E. Karakurum. Anat. Histol. Embryol, 33:278–283,
2004.
[50] Emma M L Chung, James P Hague, and David H Evans. Physics in medicine and biology,
52(23):7153–66, December 2007.
[51] JP Hague and EML Chung. Physical Review E, 80(5):051912, 2009.
[52] J.P. Hague, C. Banahan, and E.M.L. Chung. Phys. Med. Bio., 58:4581, 2013.
[53] E.M.L. Chung et al. PLOS one, 10:e0122166, 2015.
[54] Y Jiang, Z Zhuang, A J Sinusas, and X Papademetris. Conference on Computer Vision and
Pattern Recognition Workshops IEEE Computer Society Conference on Computer Vision and
Pattern Recognition. Workshops, pages 178–185, June 2010.
25
[55] Pieter Bruyninckx, Dirk Loeckx, Dirk Vandermeulen, and Paul Suetens. Segmentation of liver
portal veins by global optimization. Imaging, 7624:76241Z–76241Z–12, 2010.
[56] E. C. Novosel, C. Kleinhans, and P. J. Kluger. Advanced Drug Delivery Reviews, 63:300–311,
2011.
[57] Sudong Kim, Hyunjae Lee, Minhwan Chung, and Noo Li Jeon. Engineering of functional,
perfusable 3d microvascular networks on a chip. Lab Chip, 13:1489–1500, 2013.
[58] Jan D. Baranski, Ritika R. Chaturvedi, Kelly R. Stevens, Jeroen Eyckmans, Brian Carvalho,
Ricardo D. Solorzano, Michael T. Yang, Jordan S. Miller, Sangeeta N. Bhatia, and Christo-
pher S. Chen. Geometric control of vascular networks to enhance engineered tissue integration
and function. Proceedings of the National Academy of Sciences, 110(19):7586–7591, 2013.
26
(a)
(b)
(c)
(d)
FIG. 8: (a) Diameter as a function of Order Number for trees with 1000 vessels. Decreasing
mb, which describes the relative energy cost of an amount of blood and the power required
to pump it, has little effect on the agreement of the diameters with morphological data.
(b) For lengths however there is an obvious effect in the larger arteries, with regimes of
high pumping cost being more accurate. The primary optimisation for high pumping cost
then is to increase the length of the largest arteries. (c) and (d) The main effect is a
change in the asymmetry of the branching of the largest arteries - for large mb, the
branches are more symmetric than for small mb. As mb becomes very small, the limiting
behaviour is broad trunks that wind around all the tissue, with a large number of very
small offshoots that supply blood in the direct vicinity of the large vessel.
27
246810100101102103Order NumberDiameter µ m Morph Data0.1 × mbmb10 × mb246810102103104105106Order NumberLength µ m Morph Data0.1 × mbmb10 × mb24681000.10.20.30.40.50.60.70.80.91Order NumberDl / Dp Morph Data0.1 × mbmb10 × mb24681000.10.20.30.40.50.60.70.80.91Order NumberDs / Dp Morph Data0.1 × mbmb10 × mbFIG. 9: Average cost and standard error vs SA steps for a 127 node tree grown in a 2D
plane.
104 Steps
106 Steps
108 Steps
5 × 109
FIG. 10: Example of trees grown for various different numbers of SA step numbers.
28
10210410610810102.72.82.933.13.23.33.4x 10−3SA StepsCost |
1509.01059 | 1 | 1509 | 2015-09-03T12:42:52 | Multiple folding pathways of proteins with shallow knots and co-translational folding | [
"physics.bio-ph",
"q-bio.BM"
] | We study the folding process in the shallowly knotted protein MJ0366 within two variants of a structure-based model. We observe that the resulting topological pathways are much richer than identified in previous studies. In addition to the single knot-loop events, we find novel, and dominant, two-loop mechanisms. We demonstrate that folding takes place in a range of temperatures and the conditions of most successful folding are at temperatures which are higher than those required for the fastest folding. We also demonstrate that nascent conditions are more favorable to knotting than off-ribosome folding. | physics.bio-ph | physics | Multiple folding pathways of proteins with shallow knots
and co-translational folding
Mateusz Chwastyk and Marek Cieplak
Institute of Physics, Polish Academy of Sciences, Al. Lotnik´ow 32/46, 02-668 Warsaw, Poland
5
1
0
2
p
e
S
3
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
9
5
0
1
0
.
9
0
5
1
:
v
i
X
r
a
October 8, 2018
Abstract
We study the folding process in the shallowly knotted protein MJ0366 within two
variants of a structure-based model. We observe that the resulting topological pathways
are much richer than identified in previous studies. In addition to the single knot-loop
events, we find novel, and dominant, two-loop mechanisms. We demonstrate that folding
takes place in a range of temperatures and the conditions of most successful folding are
at temperatures which are higher than those required for the fastest folding. We also
demonstrate that nascent conditions are more favorable to knotting than off-ribosome
folding.
1
Introduction
Sufficiently long polymers, such as DNA, are likely to be entangled [1, 2]. On the other
hand, there are very few knotted RNA molecules -- in a recent assessment [3] only three
cases have been identified, but the corresponding structures are resolved poorly. Occur-
rence of knotted proteins is in between -- several hundreds of knot-containing structure
files and a dozen of truly independent structures [4, 5, 6, 7] in the Protein Data Bank
(PDB). Knots in proteins can be detected experimentally through stretching [8] since
their effective contour length is reduced, as analyzed theoretically in refs. [9, 10, 11, 12].
However, figuring out how knots get formed during the folding process is more challeng-
ing. New experimental techniques based on fusion proteins [13, 14] have started to offer
clues about the process. Nevertheless computer simulations are expected to offer more
detailed insights.
In this paper, we analyze pathways of folding in a model shallowly
knotted protein and reveal a rich complexity of possible behaviors. We also propose a
mechanical model of the ribosome and show that nascent conditions favor knotting but
also affect the pathways of folding.
Backbones of proteins do not form closed loops which leads to some ambiguities when
deciding about the presence of a knot. Nevertheless, it is often straightforward to identify
knot ends by observing knot's unknotting on cutting away sites from the termini [17, 18].
A knot is considered shallow if at least one of its ends is close to a terminus. Otherwise,
the knot is considered to be deep.
Deeply knotted proteins such as YibK have been studied experimentally [15, 16]. They
are known to fold with difficulty in simulations. A folding process here is considered
successful if the native contacts are established and the knot is formed properly. The
success rate, S, is, at best, 1 - 2 % [19, 20]. We have recently argued [21] that nascent
conditions [22, 23, 24, 25, 26] enhance the probability for YibK to become knotted when
formed on the ribosome. We have confirmed that the successful folding pathway goes
through a slipknot conformation, as suggested in ref. [19] for the off-ribosome situation.
1
We have shown that the process takes place only in a range of optimal temperatures, and
does not require any non-native interactions (if a proper procedure for their selection is
adopted).
In this paper, we focus on a protein with the native shallow trefoil knot: MJ0366
from Methanocaldococcus jannaschii which is a thermophilic methanogenic archaea. This
is the smallest of the known knotted proteins and we shall refer to this protein by its
PDB structure code of 2EFV. Folding of 2EFV has been studied theoretically [27, 28]
at a fixed temperature (T ). The studies involved all-atom simulations. Ref.
[27] used
a simplifying implicit solvent approach combined with a bias implemented through the
dominant reaction pathway method. Out of 32 successful trajectories, 26 involved direct
threading (DT), 3 -- slipknotting (SK), and 2 -- mousetrapping (MT) as mechanisms
of knotting. These mechanisms are illustrated in Figure 1. On the other hand, the
simulations in ref.
[28] took a slipknot conformation as the initial state of the system
without exploring other possible pathways: out of 15 40-µs-simulations, 5 resulted in
correct folding.
Here, we use two variants of a structure-based coarse-grained model, in which the
protein is represented as a chain of beads located at the α-C positions, and consider
various T 's and much larger statistics of between 100 to 300 trajectories for each T . We
show that: a) 2EFV gets to the knotted native state much easier than the deeply knotted
proteins, b) 2EFV should fold through a qualitatively richer family of pathways than
considered in ref.
[27], c) there is a range of optimal T 's for successful knotting, d) the
T -range corresponding to the fastest folding is shifted downward relative to the T -range
of the optimal knotting, e) nascent conditions boost the peak success rate to fold in one
of the variants of the model (in the other variant the off-ribosome peak success rate is
already 100%), and f) nascent conditions reduce the set of knotting pathways.
All of the knotting mechanisms identified in ref.
[27] are topologically single-stage
processes in which just one knot-loop is formed. We find that about 40% of our successful
trajectories indeed belong to this class. However, the majority of the trajectories involve
two stages and two smaller knot-loops. Each of the stages makes use of variants of the
DT, and SK events, but we also identify one more - an "embracement" (EM). It is possible
that the multiple-loop mechanisms of knotting are also relevant for the homopolymer-like
DNA, but have not been identified yet.
2 Structure-based modeling
The details of our approach are described in refs. [29, 30, 31]. The model is Go-like [32] so
that the length-related parameters in the potentials are derived from the native structure.
The molecular dynamics employed deals only with the α-C atoms. The bonded interac-
tions are described by the harmonic potentials. Non-bonded interactions, or contacts, are
assigned to pairs of amino acids by using the overlap criterion in which the heavy atoms
in the native conformation are represented by enlarged van der Waals spheres [29, 33]: if
at least two such spheres from different residues overlap we declare existence of a native
contact. These contacts are described by potentials with the minima at the crystallo-
graphically determined distances. The potentials are identical in depth, denoted as ǫ.
Non-native contacts are considered repulsive.
In order to test the robustness of our results, we consider two variants of the model: C
and A. In model C, the contact potentials are of the Lennard-Jones form and the backbone
stiffness is accounted for by the chirality potential [29] which favors the native sense of the
local backbone chirality. The value of ǫ has been calibrated by making comparisons to the
experimental data on stretching: approximately, ǫ/A is 110 pN (which also is close to the
energy of the O-H-N hydrogen bond of 1.65 kcal/mol). In model A, the contact potentials
are of the 10-12 kind and the backbone stiffness is described by the more common bond
and dihedral angles with the parameters specified in ref.
[34]. Model C does not have
2
the bond angle part of model A [30] and in model A, there are no i, i + 3 contacts.
The simulations are done at various temperatures. For most unknotted proteins, optimal
folding takes place around T = 0.3 ǫ/kB in model C (see also ref.
[35]) and around
T = 0.6 ǫ/kB in model A (kB is the Boltzmann constant; the stiffness parameters depend
on ǫ). Both characteristic values of T should correspond to a vicinity of the room T in
the respective models.
We use the Langevin thermostat with substantial damping. The time unit of the
simulations, τ , is effectively of order 1 ns as the motion of the atoms is dominated by
diffusion instead of being ballistic. Folding is usually declared when all native contacts
are established for the first time (the distance between two α-C in a contact is smaller
than the native distance multiplied by 1.5). For knotted proteins, however, this condition
does not necessarily signify that the correct native knot has been formed. The situation
in which there is no knot but all contacts are established is referred to as misfolding.
3 Folding of 2EFV
Protein 2EFV comprises 87 residues but the atomic coordinates of the first five of them
are not provided in the structure file. In ref. [28], the authors extend the C-terminus (i.e.
not where the residues are missing) by a 5-piece helical segment to enhance the definition
of the starting slipknot conformation. However, we find this procedure to deteriorate
folding properties so we show only the results obtained without such an extension. The
secondary of 2EFV consists of 4 helices (23-32, 41-49, 62-71, 74-86) 2 3-10 helices (33-35,
72-73), and 2 β-strand (12-17, 54-59). The knot ends in 2EFV are located at 11 and 73.
Figure 2 summarises the properties of 2EFV in model C. The inset in the lower panel
refers to the equilibrium quantities. It shows the probability, P0, of all native contacts
being simultaneously established as a function of T . Similar to the lattice models of
proteins [36] (see also an exact analysis [37]), one may define Tf as a temperature at
which P0 crosses through 1
2 -- it is 0.22 ǫ/kB in this model. The T at which the fraction
of the established native contacts, Q, crosses through 1
2 is higher, 0.75 ǫ/kB, as it signifies
the on-cooling onset of globular conformations. This T will be denoted as TQ.
The top panel of Figure 2 shows the percentage-wise success, S, of reaching the prop-
erly knotted folded conformation and the corresponding median folding time tf as a
function of T . The median times have been determined only within the subset of tra-
jectories that resulted in folding. The starting conformations are nearly fully extended.
Folding is seen to take place fast in the T -range between 0.2 and 0.55 ǫ/kB. However, the
majority of the trajectories result in knotting only between 0.45 and 0.5 ǫ/kB where P0
is 0. Thus the peak success rates involve folding times that are longer than optimal. At
T = 0.3 ǫ/kB, S is about 5% which is still better than the peak success rates reported for
the deeply knotted proteins. The bottom panel shows the S corresponding to the mis-
folding events. In the T -range corresponding to the optimal tf 's most of the trajectories
result in misfolding. S for correct folding and S for misfolding do not add up to 100% as
a portion of the trajectories does not establish all native contacts within a cutoff time of
1 000 000 τ .
Figure 3 summarises the properties of 2EFV in model A. The characteristic tempera-
tures that relate to the kinetics move upward by some 0.3 ǫ/kB and Tf shifts to 0.4 ǫ/kB
while TQ shifts to 1.0 ǫ/kB. Even though the peak S for folding is achieved just at the
upper edge of the kinetic optimality (at T = 0.9ǫ/kB ), S is larger than 50 % in the whole
range of the optimal kinetics (from 0.75 to 0.9 ǫ/kB. The peak value of S for misfolding
(68 %) is at T = 0.6ǫ/kB and it disappears at 0.9 ǫ/kB in a gradual way.
Interestingly, we find that attachment of the 5-, 10-, 15-, and 20-residue N-terminal
extensions destroys the proper folding completely -- we would expect that all-atom folding
with the extensions should be even harder because of the many more degrees of freedom
that need to cooperate. However ref.
[28] states the opposite. The extensions that we
3
have considered are either random alanine segments or a 5-residue helical extension (88-
SER, 89-ALA, 90-ASN, 91-LEU, 92-LEU). The misfolding events, on the other hand, are
observed to be frequent.
We observe that shallowly knotted proteins fold and knot much easier than the deeply
knotted ones -- but what are the mechanisms involved? Figure 4 shows that the proper
pathways fall into two classes: with a single knot-loop (the left hand side of the figure)
formed from the segment between sites 16 -- 78 or with two knot-loops (the right hand side
for the figure) between sites 16 -- 53 (shown in red) and 53 -- 78 (in blue). The two classes
arise in both models. In model C, the two-loop mechanisms occur with the T -averaged
probability of 58% and in model A -- 63%. In the single-loop class, the topological events
involve the C-terminal parts. In the DT mechanisms, the C-terminus threads through
the knot-loop (45 % in model C). In the SK mechanism, the C-terminal slipknot slides
through the knot-loop (the mechanism invoked for the deeply knotted proteins; 45 %).
In the MT mechanism, the knot-loop moves to envelop the C-terminus (10 %). A typical
time scale in which the first loop forms (also in the single-loop mechanism) is around
4 000 τ in both models. It takes much longer to form the second loop -- typically 100 000
τ more. At high T 's (like T = 1.0 ǫ/kB) in model A, it takes of order 500 000 τ to form
a globular state followed a sudden single-loop knotting.
In the two-loop class of pathways, two smaller knot-loops are formed and the events
involve both termini. Typically, the topological transformations start at the N-terminus
and, at stage A, split into three pathways. They correspond to the mechanisms of DT (6
%), SK (38 %) and EM (56 %). The latter is one in which segment 53-78 "embraces" the
mostly idle N-terminal part and forms the knot-loop around it. The next stage, denoted
as B, knotting is completed by engaging the C-terminal segment. Here, the pathways split
into four mechanisms: SK (44%), MT (11%), DT (39%), and EM (6%). There are some
events in which stages A and B are interchanged. An example of a situation in which one
EM follows another EM mechanism is shown in Figure 5.
In ref.
[27] the knotting mechanisms have been apparently classified based on the
events that shortly precede the appearance of knots without identifying the one- and two-
loop pathways. If we cumulate the four IIB events with the I events and take the weighted
average between the two classes of pathways, we get about 44% in SK, 41% in DT, 11%
in MT, and 4% in EM modes which does not agree with about 81% in DT obtained in
ref. [27]. This could be due to either the role of the side groups, or the choice of just one
T which may not be optimal, or the differences in the statistics.
We now consider a start from the slipknotted conformations without the extensions.
If we start from SK - IIB, we get a 96% success rate at T = 0.35ǫ/kB and 98% at
T = 0.45 ǫ/kB. On the other hand, if we starts from SK - IIA then the success rates are
correspondingly 34% and 96%. Finally, if we start from SK - I, similar to ref.
[28], we
get 97 % at T = 0.45 ǫ/kB. Our success rate is three times higher than in the all-atom
simulations, which may reflect the Go-like character of our model and the lack of the
water molecules. However Noel et al. [28] extend the protein at the C-terminus which we
find to deteriorate folding.
4 Folding of 2EFV under nascent conditions
The percentage-wise success, S, of reaching the properly knotted folded conformation in
model C increases substantially when simulating the process in the co-translational way.
In our previous model of on-ribosome folding [21] we have focused on the most essential
aspects: the excluded volume provided by the ribosome and the related reduction in the
conformational entropy are captured by representing the ribosome as an infinite plate
which spawns a protein residue by residue at one fixed location. We take the plate to
generate a laterally uniform potential of the form 3√3
z )9, where z denotes the distance
1/6. The proteins are synthesized from the N terminus
away from the plate and σ0 = 4×2−
ǫ ( σ0
2
4
to the C terminus. The time interval between the emergence of two successive α-C atoms,
tw, is taken as 5000 τ since larger values lead to saturation in S. Once the backbone is
formed fully, the protein is released and then evolved up to a cutoff time of 1 000 000
τ . Our model is illustrated by Figure 6 which shows the protein toward the end of the
process in which it is being born.
Figure 7 shows that the nascent conditions enhance the probability of establishing the
native contacts and the peak success rate of knot formation. They also change the look of
the dependence of S for knotting on T . For the deeply knotted protein the enhancement
in S made a qualitative difference as it enabled folding [21]. For the shallowly knotted
2EFV, the kinetic improvement is minor -- by 5 percentage points from the peak value
of S=76 %. However, we observe a significant shift in the occurrence of the topological
events: the single-loop pathways disappear and the role of the SK events, so crucial for
the deeply knotted proteins, gets diminished substantially. Specifically, in step IIA shown
in the right hand side of Figure 4 the DT events disappear, the SK events are reduced to
5% and the EM mechanism is operational in 95% of the successful trajectories. In stage
IIB, MT and DT are enhanced (to 20 and 55% respectively) whereas SK gets diminished
(25%) and EM disappears. The nascent conditions favor loops that first form at the
N-terminus.
5 Conclusions
Our results for shallowly knotted 2EFV indicate easy and correct folding with the topolog-
ical optimality shifted upward in T relative to the kinetic optimality. We have identified
novel two-loop mechanisms of knotting and showed that the topological mechanisms are
much richer than discussed before. Their variants may also be operational in homopoly-
mers. In our studies of YibK, we have found that extending the overlap-based contact
map by extra contacts identified by the CSU server [38] was vital for the success of on-
ribosome folding. However, for 2EFV adding the extra contacts is found not to affect
any of the results discussed here. On-ribosome folding is more efficient than off-ribosome
and is also more selective in its mechanisms. It would be interesting to consider a real-
istic variant of co-translational folding -- one that takes into account confinement. This
is because the ribosome spawns proteins into molecularly sculpted chambers [39], which
parallels the physics of encapsulation within GroEL-GroES chaperonins considered by
Lim and Jackson [14] in the context of deeply knotted proteins.
Acknowledgments We appreciate useful correspondence with C. Micheletti and J.
Trylska. This work has been supported by the National Science Center in Poland under
the aegis of the EU Joint Programme in Neurodegenerative Diseases (JPND). The local
computer resources were financed by the European Regional Development Fund under
the Operational Programme Innovative Economy NanoFun POIG.02.02.00-00-025/09.
References
[1] J. M. Arsuaga, S. Vazquez, S. Trigueros, D. W. Sumners, J. Roca, Proc. Natl. Acad.
Sci. USA 99, 5373-5377 (2002).
[2] D. Marenduzzo, C. Micheletti, E. Orlandini, Physics World 30-34 April, (2013).
[3] C. Micheletti, M. Di Stefano, H. Orland, Proc. Natl. Acad. Sci. USA 112, 2052-2057
(2015).
[4] P. Virnau, L.A. Mirny, M. Kardar, Plos. Comp. Biol. 2, e122 (2006).
[5] P. Virnau, A. Mallam, S. Jackson, J. Phys. Cond. Mat. 23, 033101 (2011).
[6] J. I. Su lkowska, E. J. Rawdon, K. C. Millet, J. N. Onuchic, A. Stasiak, Proc. Natl.
Acad. Sci. USA 109, E1715-E1723 (2012).
5
[7] M. Jamroz, W. Niemyska, E. J. Rawdon, A. Stasiak, K. C. Millet, P. Su lkowski, J.
I. Su lkowska, Nucl. Acid. Res. 43, D306-314 (2015).
[8] D. M. Anstrom, E. Mey, J. Dziubiella, M. Rief, K. T. Forest, Biophys. J. 96, 1508-
1514 (2009).
[9] J.I. Su lkowska, P. Su lkowski, P. Szymczak, M. Cieplak, Phys. Rev. Lett. 100, 058106
(2008).
[10] J.I. Su lkowska, P. Su lkowski, P. Szymczak, M. Cieplak, J. Am. Chem. Soc. 132,
13954-13956 (2010).
[11] J. Dziubiella, Biophys. J. 96, 831-839 (2009).
[12] M. Chwastyk, M. Cieplak, Israel Journal of Chemistry 54, 1241-1249 (2014).
[13] A. L. Mallam, S. C. Onucha, J. G. Grossman, S. E. Jackson, Mol. Cell. 30, 642-648
(2008).
[14] N. C. H. Lim, S. E. Jackson, J. Mol. Biol. 427, 248-258 (2015).
[15] A. L. Mallam, S. E. Jackson, Nat. Chem. Biol. 8, 147-153 (2012).
[16] A. L. Mallam, J. M. Rogers, S. E. Jackson, Proc. Natl. Acad. Sci. USA 107, 8189-
8194 (2010).
[17] K. Koniaris, M. Muthukumar, Phys. Rev. Lett. 66, 2211-2214 (1991).
[18] W. Taylor, Nature 406, 916-919 (2000).
[19] J. I. Su lkowska, P. Su lkowski, J. N. Onuchic, Proc. Natl. Acad. Sci. USA 106,
3119-3124 (2009).
[20] W. Li, T. Terakawa, W. Wang, S. Takada, Proc. Natl. Acad. Sci. USA 109, 17789-
17794 (2012).
[21] M. Chwastyk, M. Cieplak, Cotranslational folding of deeply knotted proteins J.
Phys. Cond. Mat. 27, 354105 (2015).
[22] L. D. Cabrita, C. M. Dobson, J. Christodoulou, Curr. Op. Struct. Biol. 20, 33-45
(2010).
[23] C. M. Kaiser, D. H. Goldman, J. D. Chodera, I. Tinoco Jr., C. Bustamante, Science
334, 1723-1727 (2011).
[24] S. Melnikov, A. Ben-Shem, N. Garreau de Loubresse, L. Jenner, G. Yusupova, M.
Yusupov, Nat. Struct. Mol. Biol. 19, 560-567 (2012).
[25] N. Garreau de Loubresse, I. Prokhorova, W. Holtkamp, M. V. Rodnina, G.
Yusupova, M. Yusupov, Nature 513, 517-522 (2014).
[26] J. D. Puglisi, Science 348, 399-400 (2015).
[27] S. a Beccara, T. Skrbic, R. Covino, C. Micheletti, P. Faccioli, PLOS Comp. Biol. 9,
e1003002 (2013).
[28] J. K. Noel, J. N. Onuchci, J. I. Su lkowska, Phys. Chem. Lett. 4, 3570-3573 (2013).
[29] J.I. Su lkowska and M. Cieplak, J. Phys.: Cond. Mat. 19, 283201 (2007).
[30] J. I. Su lkowska, M. Cieplak, Biophys. J. 95, 3174-3191 (2008).
[31] M. Sikora, J.I. Su lkowska, M. Cieplak, PLOS Comp. Biol. 5, e1000547 (2009).
[32] N. Go, Annu. Rev. Biophys. Bioeng. 12, 183-210 (1983).
[33] J. Tsai, R. Taylor, C. Chothia, M. Gerstein, J. Mol. Biol. 290, 253-266 (1999).
[34] C. Clementi, H. Nymeyer, J.N. Onuchic, J. Mol. Biol. 298, 937-953 (2000).
[35] M. Cieplak, T. X. Hoang, Biophys. J. 84, 475-488 (2003).
6
[36] N. D. Socci, J. N. Onuchic, J. Chem. Phys. 101, 1519-1528 (1994).
[37] M. Cieplak and J. R. Banavar, Phys. Rev. E. Rapid Comm. 88, 040702(R) (2013).
[38] V. Sobolev, A. Sorokine, J. Prilusky, E.E. Abola, M. Edelman, Bioinformatics 15,
327-332 (1999).
[39] A. H. Elcock, PLOS Comp. Biol. 2, e98 (2006).
7
SK
SKDT
DTMT
Figure 1: The folding mechanisms in 2EFV found in ref.
[27]. SK denotes folding through
slipknotting, DT -- through direct threading, and MT -- through the mousetrap-like mechanism.
Each process is illustrated by showing two subsequent stages. SK involves sliding of a slipknot
through the knot-loop. In DT, the terminus threads through the knot-loop. MT is similar to
DT but the knot-loop makes the dominant movement instead of the terminal site. The orange
segment extends from the terminal N to site 16, the red segment -- from 17 to 53, the blue
segment -- from 54 to 78, and the gray segment is the remaining C-terminal piece.
8
Figure 2: Properties of 2EFV in model C. The inset in the lower panels shows P0 and Q as a
function of T . The main panels characterize the kinetic quantities: the upper panels are for
correct folding and the lower panels -- for misfolding. The open squares correspond to S -- the
success rate. The solid circles correspond to the median folding times. These data points are
obtained by starting from extended conformations. The data points denoted by the hexagons
and asterisks correspond the the slipknotted conformations shown in Figure 1.
Figure 3: Similar to Figure 2 but for model A.
9
Figure 4: Stages of folding and knotting process of protein 2EFV. The top panel shows the
extended conformation which is divided into 4 sequential segments: the N-terminal part is in
yellow, C-teminal part is in gray, the first knot-loop is in red, and the second knot loop in
blue. The bottom panel gives a schematic representation of the final native state. Block I,
on the left, shows the three single-loop mechanisms of knotting: DT -- direct threading, SK --
slipknotting, MT -- mousetrapping. Block II, on the right, shows two stages, A and B, of the
two-loop mechanisms of knotting. In addition to DT, SK, and MT, they also involve EM --
embracement. The smaller loops have radii between 8 and 10 A.
10
Figure 5: An example of a two-loop mechanisms of folding combined with knotting. Here the
embracement of the N-terminus (the top panel) is followed by a similar event at the C-terminus
(the bottom panel). The panels on the left show the protein's conformation at the beginning
of the knotting at the particular stage. The right panels show the resulting state.
16
N
53
75
Figure 6: Nascent protein 2EFV emerging from the model ribosome. At the stage shown, the
segment 75 -- C is not yet born. The colors of the segments in the backbone approximate those
of Figure 4.
11
Figure 7: The upper panel shows the success rates for the on-ribosome (the solid squares) and
off-ribosome (the open squares) folding as a function of T . The median folding times, tf , are
written for T = 0.4 ǫ/kB (13000 τ is for the off-ribosome folding). The lower panel is similar
but shows the data for misfolding.
12
|
1509.07219 | 3 | 1509 | 2016-02-12T23:26:34 | Motor protein accumulation on antiparallel microtubule overlaps | [
"physics.bio-ph",
"q-bio.SC"
] | Biopolymers serve as one-dimensional tracks on which motor proteins move to perform their biological roles. Motor protein phenomena have inspired theoretical models of one-dimensional transport, crowding, and jamming. Experiments studying the motion of Xklp1 motors on reconstituted antiparallel microtubule overlaps demonstrated that motors recruited to the overlap walk toward the plus end of individual microtubules and frequently switch between filaments. We study a model of this system that couples the totally asymmetric simple exclusion process (TASEP) for motor motion with switches between antiparallel filaments and binding kinetics. We determine steady-state motor density profiles for fixed-length overlaps using exact and approximate solutions of the continuum differential equations and compare to kinetic Monte Carlo simulations. Overlap motor density profiles and motor trajectories resemble experimental measurements. The phase diagram of the model is similar to the single-filament case for low switching rate, while for high switching rate we find a new low density-high density-low density-high density phase. The overlap center region, far from the overlap ends, has a constant motor density as one would naively expect. However, rather than following a simple binding equilibrium, the center motor density depends on total overlap length, motor speed, and motor switching rate. The size of the crowded boundary layer near the overlap ends is also dependent on the overlap length and switching rate in addition to the motor speed and bulk concentration. The antiparallel microtubule overlap geometry may offer a previously unrecognized mechanism for biological regulation of protein concentration and consequent activity. | physics.bio-ph | physics |
Motor protein accumulation on antiparallel microtubule overlaps
H.-S. Kuan and M. D. Betterton1
Department of Physics,
University of Colorado Boulder, Boulder, CO
1Corresponding author. Address: Department of Physics, University of Colorado Boulder 390 UCB
Boulder, CO 803095, U.S.A., Tel.: (303)735-6135
Abstract
Biopolymers serve as one-dimensional tracks on which motor proteins move to perform their biolog-
ical roles. Motor protein phenomena have inspired theoretical models of one-dimensional transport,
crowding, and jamming. Experiments studying the motion of Xklp1 motors on reconstituted an-
tiparallel microtubule overlaps demonstrated that motors recruited to the overlap walk toward the
plus end of individual microtubules and frequently switch between filaments. We study a model
of this system that couples the totally asymmetric simple exclusion process (TASEP) for motor
motion with switches between antiparallel filaments and binding kinetics. We determine steady-
state motor density profiles for fixed-length overlaps using exact and approximate solutions of the
continuum differential equations and compare to kinetic Monte Carlo simulations. Overlap motor
density profiles and motor trajectories resemble experimental measurements. The phase diagram
of the model is similar to the single-filament case for low switching rate, while for high switching
rate we find a new low density-high density-low density-high density phase. The overlap center
region, far from the overlap ends, has a constant motor density as one would naıvely expect. How-
ever, rather than following a simple binding equilibrium, the center motor density depends on
total overlap length, motor speed, and motor switching rate. The size of the crowded boundary
layer near the overlap ends is also dependent on the overlap length and switching rate in addition
to the motor speed and bulk concentration. The antiparallel microtubule overlap geometry may
offer a previously unrecognized mechanism for biological regulation of protein concentration and
consequent activity.
Key words: motor proteins; microtubules; TASEP; cytoskeleton
Motors on antiparallel MT overlaps
2
Introduction
The motion of motor proteins on biopolymers is important for diverse biological processes (1).
Actin, microtubules, and nucleic acids can serve as one-dimensional tracks on which motor proteins
(including myosins, kinesins, helicases, and ribosomes) move (2, 3). Motors must accumulate on
filaments in sufficient density to perform their biological roles. This motor accumulation along a
biopolymer is affected by two key effects: the directional walking of motors and motor binding
to/unbinding from the filament.
Motor accumulation on filaments is related to extensive theoretical work on asymmetric exclu-
sion processes (ASEP) (4). In ASEP models, particles move on a one-dimensional (1D) lattice by
biased hopping and experience excluded-volume interactions with other particles. These models
have been applied to diverse examples of one-dimensional nonequilibrium transport ranging from
molecular motors to vehicular and pedestrian traffic. Active particle motion leads to non-zero flux
of particle density and nontrivial flux and density profiles. To compare to experimental studies of
molecular motors, ASEP models have been extended to incorporate important biophysical ingredi-
ents such as binding kinetics. A model extending the totally asymmetric exclusion process (TASEP)
to include motor binding and unbinding (Langmuir kinetics) predicted motor density profiles along
single fixed-length filaments (5). Experimental work measured kinesin-8 motor protein traffic jams
on microtubules and found good agreement with the predicted density profiles (6).
Motor motion on cytoskeletal filaments is important for biological length regulation, including
regulation of the length of the polymer (7), lengths of overlap regions between filaments (8), and
the length of cytoskeletal assemblies such as the mitotic spindle (9, 10) and even whole cells (11,
12). The case of regulation of microtubule (MT) length has seen the most work. MTs undergo
nonequilibrium polymerization dynamics characterized by switching between distinct growing and
shrinking states. While this dynamic instability alone leads to a broad distribution of MT lengths
(13), numerous proteins targeted to MT ends modify their dynamics and can dramatically alter
the length distribution (14). Single MTs can have their length regulated, for example, by kinesin-8
motors that walk with directional bias and shorten the MT from its end (7, 15, 16). Theoretical
work on TASEP-like models has described how length-dependent depolymerization affects otherwise
static filaments (7, 17, 18), filaments with simplified polymerization kinetics, (19 -- 22), and dynamic
MTs (23 -- 25).
Because the mitotic spindle includes arrays of overlapping antiparallel MTs at the spindle
midzone, regulation of MT overlaps is important for mitotic spindle function and cytokinesis. The
MT crosslinking protein PRC1/Ase1/MAP65 and kinesin-4 motors (chromokinesins) play roles in
maintenance of the spindle midzone in anaphase (26 -- 28), along with other motors and MAPs (29).
Direct binding interactions of PRC1 and kinesin 4 can cause length-dependent accumulation at
plus ends of single MTs (30). Bieling, Telley, and Surrey (BTS) reconstituted a minimal system
of stable antiparallel MT overlaps in which PRC1 bound preferentially to overlapping regions of
antiparallel MTs (8) (see also Subramanian et al. (31)). PRC1 recruited the kinesin-4 motor Xklp1
to the overlap. Xklp1 motors could bind to and unbind from the MTs, walk toward the plus end of
each MT, and switch between the two MTs at a relatively high rate (8). Motors present near the
MT plus ends slowed the polymerization speed, consistent with earlier work showing that Xklp1
inhibits dynamic instability (32) and affects spindle MT mass (33). As a result, antiparallel MT
overlaps reached a constant length that depended on the bulk concentration of motors. This work
demonstrated that motor-dependent regulation of dynamics and length can occur not just for single
MTs, but for overlapping MT pairs.
Here we model the motor density profiles of motor proteins on antiparallel MT overlaps of fixed
lengths. We do not explicitly consider MT length regulation in this system, as the combination
Motors on antiparallel MT overlaps
3
of binding kinetics and coupled switching of the motors between the two antiparallel filaments is
sufficient to produce a rich phenomenology. Our work is an extension of Parmeggiani, Franosch,
and Frey's TASEP with binding kinetics on a single filament (5, 34) to include two antiparallel
filaments coupled by switching; it is also an extension of TASEP models of two antiparallel lanes
with lane switching (35, 36) to incorporate binding and unbinding kinetics.
We first develop the model and show that measured and estimated parameters can give overlap
motor density profiles and motor trajectories qualitatively similar to those found in the BTS ex-
periments. We then develop an analytic solution to the mean-field steady state equations, and use
it to determine the phase diagram of the model. For high motor switching rate between filaments,
we find a new low density-high density-low density-high density phase. We then study the model
for the reference parameter set in more detail. Because the motor density profiles are controlled
by both boundary conditions and a non-local total binding constraint, the density in the center
of the overlap depends not just on motor binding kinetics but also on motor speed and overlap
length. We find an analytical approximation that describes the overlap center density and the size
of the motor-dense boundary layer near the overlap ends. The degree of motor accumulation near
the overlap ends depends on the overlap length and motor switching rate in addition to the motor
speed and bulk concentration. The coupling of motor motion, binding, and switching kinetics on
antiparallel MT overlaps may therefore offer a previously unrecognized mechanism for the control
of one-dimensional motor density profiles.
Materials and Methods
Model
In this section, we develop a mathematical model of motor density on antiparallel filament overlaps,
inspired by the experiments of Bieling, Telley, and Surrey (BTS) (8). In the BTS experiments,
the crosslinking protein PRC1 binds preferentially to overlapping regions of antiparallel MTs and
recruits the kinesin motor protein Xklp1 to the overlap region. Because PRC1 and therefore
the motors are present at much higher concentrations in the overlap, in the model we consider
the overlap region only and consider the regions of single filaments as sources or sinks of motor
proteins (fig. 1A). While previous work has shown that PRC1/Ase1 alone can develop density
inhomogeneities and exert forces on sliding MTs (37, 38), in the BTS experiments little to no MT
sliding occurred and the PRC1 distribution was uniform in the overlaps (8). Therefore we do not
explicitly model the PRC1 molecules or their spatial distribution, but assume they are uniformly
distributed so that each lattice site in the overlap is identical with a binding affinity that would
correspond to the density-weighted average of the PRC1 and bare tubulin affinities. We treat each
MT as a single track and neglect the multiple-protofilament structure of the MT, consistent with
previous theoretical work (7, 17 -- 22, 24). In our model, motors can bind to and unbind from each
of the MTs, walk toward the plus end of a MT, and switch between the two MTs (fig. 1A).
There are two competing processes in this model: motor stepping (TASEP) and motor binding,
unbinding, and MT switching (Langmuir kinetics). The Langmuir kinetics dominate the system
behavior if the overlap length is sufficiently large, N (cid:29) 1, so that each bound motor can only
walk for a short fraction of the overlap length before it unbinds. Thus, the competition between
TASEP and Langmuir kinetics occurs only if the overall binding rate to one MT in the overlap
Konc = N konc (where c is the bulk motor concentration), and unbinding rate Koff = N koff , are of
similar magnitude to the motor speed v (5).
In the next section, we develop the discrete microscopic model and corresponding stochastic
simulation, then derive the mean-field continuum description.
Motors on antiparallel MT overlaps
4
Discrete model
We consider two overlapping antiparallel MTs of fixed length, so the number of sites N is fixed
(fig. 1A). At each site, motor binding or unbinding can occur with binding rate konc, where kon
is the binding rate constant per site and c the bulk motor concentration, and unbinding rate koff .
Each bound motor steps at rate v to the next site toward the MT plus end (if the next site is
unoccupied), and switches at rate s to the site on the adjacent MT (if that site is unoccupied).
Note that Xklp1 is plus-end directed, and we therefore assume that the motion on a single MT is
unidirectional.
for interior sites (2 < i < N − 1) on MTs with plus end to the right (R) and left (L) are
The occupation number ni is 1 if site i is occupied or 0 if site i is empty. Then the equations
dnR,i(t)
dt
dnL,i(t)
dt
=
=
vnR,i−1(t)[1 − nR,i(t)] − vnR,i(t)[1 − nR,i+1(t)] + konc[1 − nR,i(t)]
−koff nR,i(t) − snR,i(t)[1 − nL,i(t)] + snL,i(t)[1 − nR,i(t)],
vnL,i+1(t)[1 − nL,i(t)] − vnL,i(t)[1 − nL,i−1(t)] + konc[1 − nL,i(t)]
−koff nL,i(t) − snL,i(t)[1 − nR,i(t)] + snR,i(t)[1 − nL,i(t)].
(1)
(2)
At the boundary sites, we modify these equations to incorporate fluxes into and out of the overlap.
The flux into the overlap is vα[1 − n1(t)], where nonzero α results from motors moving into the
overlap from the adjacent single-MT region. The flux out of the overlap is vβ nN (t), where β is
derived fromthe rate at which motors at an MT plus end unbind. In principle, each filament could
have different boundary conditions. Because we have no physical reason to distinguish the two
halves of the overlap, we focus on the symmetric case αL = αR = α and βL = βR = β. The
resulting boundary site equations are
dnR,1(t)
dt
dnR,N (t)
dt
dnL,1(t)
dt
dnL,N (t)
dt
= vα[1 − nR,1(t)] − vnR,1(t)[1 − nR,2(t)],
= vnR,N−1(t)[1 − nR,N (t)] − vβ nR,N (t),
= vnL,2(t)[1 − nL,1(t)] − vβ nL,1(t),
= vα[1 − nL,N (t)] − vnL,N (t)[1 − nL,N−1(t)].
(3)
(4)
(5)
(6)
As shown below, these boundary conditions fix the motor densities to be α at the minus end and
1 − β at the plus end of each filament.
At steady state, we can derive a total binding constraint on the equations by summing all of
sites on both filaments. The bulk flux terms of the form ni−1(t)[1− ni(t)] sum to zero and only the
binding and boundary terms remain:
[2konc − (konc + koff )(nR,i + nL,i)] + vα[2 − nR,1(t) − nL,N (t)] − vβ[nR,N (t) + nL,1(t)] = 0. (7)
N−1(cid:88)
i=1
i=2
N(cid:88)
We find a binding constraint on the total motor binding
nR,i + nL,i = 2N
konc
konc + koff
+
2v[α(1 − α) − β(1 − β)]
konc + koff
= 2N ρ0 +
2v[α(1 − α) − β(1 − β)]
konc + koff
, (8)
Motors on antiparallel MT overlaps
5
where we have defined the Langmuir density ρ0 = konc/(konc + koff ). Therefore, at steady state
an equilibrium involving binding, unbinding, and the filament-end boundary conditions must be
reached on average for the entire overlap. This is related to the zero-current condition found in
previous work on the antiparallel TASEP without binding kinetics (35, 36). We did not find an
analytical solution to the discrete equations. Instead, we performed kinetic Monte Carlo (kMC)
simulations of the discrete model (Supporting Material) and compared to solutions in the continuum
limit.
Mean-field continuum model
We derive the mean-field continuum model as in previous work (5) by taking the stationary average
(cid:104)ni(cid:105) ≡ ρi, applying the random phase approximation (cid:104)nini+1(cid:105) = (cid:104)ni(cid:105)(cid:104)ni+1(cid:105). We also assume
motor commutation during track switching (cid:104)nR,inL,i(cid:105) = (cid:104)nL,i(cid:105)(cid:104)nR,i(cid:105) to give discrete mean-field
equations with linear switching terms (Supporting Material). We take the continuum limit and
nondimensionalize the parameters and variables by choosing the length of the overlap, L, as the
unit of length and L/v as the unit of time. Capital letters denote the nondimensionalized parameters
(S = sL/v and so on, Supporting Material). We choose x = 0 as the center of the overlap, and
the boundary conditions become ρR(−0.5) = ρL(0.5) = α and ρR(0.5) = ρL(−0.5) = 1 − β. The
steady-state continuum mean-field equations are
0 = (2ρR − 1)
0 = (1 − 2ρL)
∂ρR
∂x
∂ρL
∂x
+ Konc(1 − ρR) − Koff ρR − SρR + SρL,
+ Konc(1 − ρL) − Koff ρL + SρR − SρL.
(9)
(10)
In the continuum mean-field model, the total binding constraint (Eqn. 8) is
dx ρL =
Konc
Konc + Koff
+
α(1 − α) − β(1 − β)
L(Konc + Koff )
= ρ0 +
α(1 − α) − β(1 − β)
L(Konc + Koff )
.
(11)
(cid:90) 1/2
−1/2
(cid:90) 1/2
−1/2
dx ρR =
Results
Motor density and trajectories in overlap
To study whether our model can qualitatively describe the motor density in antiparallel MT over-
laps observed experimentally, we determined a reference parameter set corresponding to the BTS
experiments (table 1). Most parameters were directly measured by BTS. We estimated the motor
on rate constant kon by comparison to the single-molecule imaging of low-density Xklp1 in an over-
lap (8). To estimate the minus-end boundary condition α, note that BTS found that Xklp1 binding
was greatly increased on overlaps due to recruitment by PRC1. In the BTS experiments, low ionic
strength (which favors motor-MT binding) was needed to observe significant Xklp1 binding to sin-
gle MTs outside of overlaps. Therefore, we assume that the flux of motors into the overlap from
outside is negligible, and set α = 0 in our reference parameters. The plus-end boundary condition
β is related to the motor unbinding rate at MT plus ends. While this end off rate was not directly
measured by BTS, kinesin motors typically pause at MT plus ends and have an end unbinding
rate smaller than the unbinding rate in the bulk. We therefore expect that β lies between 0 (no
end unbinding) and 2.7 × 10−3 (the value of β corresponding to an unbinding rate equal to koff ,
the motor unbinding rate in the bulk). We found that β = 2.7 × 10−3 is so small that the motor
Motors on antiparallel MT overlaps
6
Symbol Parameter
v
Motor speed
Reference value
0.5 µm s−1
kon
Binding rate constant
2.7 × 10−4 nM−1 s−1
c
koff
s
α
β
N
δ
Bulk motor concentration
Unbinding rate
Switching rate
Motor flux constant
into
overlap from MT minus end
1 -- 200 nM
0.169 s−1
0.44 s−1
0
Motor flux constant out of
overlap from MT plus end
0
Number of sites
Length of a single site
120 -- 2500
8 nm
Notes
Measured by Bieling et al. (8); varied up to
8 µm s−1 to study effect of varying speed
on density profiles
Estimated based on motor density profiles
and kymographs to give affinity of 6.3 nM,
∼15x higher than measured for motor on
single MT by Bieling et al. (39)
Varied by Bieling et al. (8)
Measured by Bieling et al. (8)
Measured by Bieling et al. (8)
Motors bind primarily inside the overlap;
see discussion in the Results. We varied α
between 0 and 1 to determine the model
phase diagram
An upper bound on the end motor unbind-
ing rate is β = 2.7 × 10−3; see discussion
in the Results. We varied β between 0 and
1 to determine the model phase diagram
Varied to study overlaps of length 1 -- 20 µm
Length of an α-β tubulin dimer, see Bray
(1)
Table 1: Parameter values for the reference parameter set, taken from experimental measurements
or estimated as noted.
density profiles were indistinguishable from those with β = 0 (fig. S1). Therefore we used β = 0 in
our reference parameter set.
With these parameters, we studied kMC simulations of our model with varying motor concen-
tration and overlap lengths corresponding to the steady-state values measured by BTS. We made
simulated images that represent how our model's motor distributions would appear in an experi-
ment with fluorescently tagged motors (green) and antiparallel MT overlap region (red) (fig. 1B,
Supporting Material). Motors decorating the overlap, with greater density and end accumula-
tion for higher bulk motor concentration. The simulated images are qualitatively similar to the
experimental images (fig. 4 in Bieling et al. (8)). We also made simulated kymographs (fig. 1C,
Supporting Material), which show directed motor motion with direction reversal, qualitatively sim-
ilar to the experimental results (figs. 6 and S5 in Bieling et al. (8)). We note that the experimental
kymographs show a larger population of paused/immobile motors than occurs in our model.
Analytic solution of the steady-state continuum equations
To further study the behavior of our model, we determined analytic steady-state solutions of the
continuum mean-field equations. One solution to Eqn. 9, 10 is the constant Langmuir density set
by binding/unbinding equilibrium, ρ0 = Konc/(Konc+Koff ). To find spatially varying solutions, we
first define σR,L = ρR,L − 1
2 , the difference of the motor occupancies from 1
2 . The rate combinations
are k = Konc + Koff + S and γ = Konc − Koff . Then the equations become
dσR
dx
dσL
dx
k
=
2
= − k
2
− γ
4σR
γ
4σL
+
− SσL
2σR
+
SσR
2σL
.
(12)
(13)
Motors on antiparallel MT overlaps
7
Eqns 12 and 13 are well defined for σR,L (cid:54)= 0. This allows solution of the differential equation
relating σR + σL and σR − σL with one integration constant (Supporting Material). Our kMC
simulation results for motor density in the overlap differ substantially from this analytic solution
(fig. S2), because the analytic solution (Eqn. S18) does not satisfy the total binding constraint
(Eqn. 11). Therefore, the full density profiles must be determined by matching continuum solu-
tions corresponding to different integration constants. This mathematical result implies several
remarkable properties of the density profiles, as discussed below.
Phase diagram
We determined the phase diagram of the model as a function of the boundary condition parameters
α and β, as well as how the phase diagram changes with the other model parameters. We find four
phases previously observed for the single-filament case (5), the low density (L), high density (H),
low density-high density (LH), and Meissner (M) phases (figs. 2, S3). We distinguish two different
classes of L, H, and LH phase on overlaps: the center-accumulating case (abbreviated c) has a
higher total motor concentration profile near the center of the filament and motor accumulation
at the overlap center, while the end-accumulating case (e) has a lower total motor concentration
profile near the filament center and motor accumulation at the overlap ends. In addition, for high
switching rate we observe a new low density-high density-low density-high density (LHLH) phase
(fig. 2).
We studied flows in the σL − σR phase plane to determine the motor concentration profiles
(fig. S4) and therefore the phase diagram, as discussed in the Supporting Material. Representative
phase diagrams for small and large switching rate are shown in fig. 3. We note that the new
LHLH phase occurs when α < 0.5 and the switching frequency is sufficiently high, conditions which
apply to the BTS experiments. This phase can appear for high switching rate when two analytic
solutions to the steady-state continuum equations intersect, causing transition points in the phase
plane where the density is ill defined (fig. S4 -- S5). Our LHLH phase is reminiscent of multi-phase
coexistence found by Pierobon (40). However, in this previous work the multi-phase co-existence
occurs due to a point defect in the lane, while here it occurs for spatially uniform dynamics.
The phase boundaries change with parameters of the model. Increases (decreases) to the Lang-
muir density (through changes in bulk motor concentration or binding/unbinding rate constants)
shift the lower boundary of the LH phase closer to (farther from) the line where 1 − β = 0.5 when
α > ρ0 and closer to (farther from) the line α = β when α < ρ0. Similar shifts occur for the upper
boundary of the LH phase. The motor speed affects the width and the shape of the LH phase:
for higher motor speed, the LH phase region becomes narrower in width and the LH boundaries
become more flat.
Density profiles for reference parameters
The motor density profiles that we determine by kMC simulation are qualitatively consistent with
those observed experimentally (figs. 1, 4). Here we illustrate our results for the reference parameter
set (table 1); in the Supporting Material we also discuss a large-S parameter set chosen to approx-
imate the limit of large switching rate (table S1, fig. S6)). This parameter regime typically shows
the LH phase, where the density profiles have three regions separated by domain walls. For motors
moving to the right, there is a boundary layer on the left with the motor density increasing from
zero, then a region of approximately linearly varying density, then a sharp transition to another
boundary layer of linearly increasing density that approaches 1 on the right boundary (fig. 4). If
we fix all other parameters, the boundary layer regions increase in size as the bulk motor concen-
Motors on antiparallel MT overlaps
8
tration increases. The total motor density in the overlap is the sum of the motor densities on the
two filaments, which have the symmetry ρL(−x) = ρR(x). The linear density variation near the
overlap ends can be seen in the continuum mean-field equations (Supporting Material): the slope
is Konc + S at the minus end and Koff + S at the plus end. This approximation agrees well with
simulation results near the overlap ends (fig. S2, S7).
Control of center density by the total binding constraint
Motor density profiles must satisfy the total binding constraint of Eqn. 11, which requires that
when α = β = 0, the integral of the density on a single MT must equal ρ0, the Langmuir density
determined by binding/unbinding equilibrium. We verified that the total binding constraint is
satisfied in our simulations by determining the integrated motor density and comparing it to ρ0
(fig. S8). However, as noted above, the analytic solutions we found for the steady-state motor
density (Eqns. S18, S23, S24) do not in general satisfy the total binding constraint. As a result,
the motor concentration in the center of the filaments does not necessarily approach the Langmuir
density ρ0 = Konc/(Konc + Koff ) (fig. 4). This is a significant difference between the model of
microtubule overlaps we study and the single-filament model (5).
The total binding constraint also controls the length of the boundary layer regions near the
overlap ends. To derive an analytic approximation expression for this length, we approximate the
motor densities as piecewise linear, and assume the domain walls are infinitely thin so that we can
neglect them in integrating the concentration (fig. S9). A filament is divided into three regions:
boundary layers near the plus and minus ends, and the central region. We can then determine the
motor density profile (Eqn. S28), the boundary layer ends, xbl (Eqn. S30), and the motor density
in the center of the overlap, ρc (Eqn. S31).
The motor density at the center of the overlap ρc (cid:54)= ρ0, even when the overlap is long. Instead
the center density depends on the motor speed and filament switching rate in addition to binding
parameters (fig. 4) In fig. 5, we compare simulation results and the prediction of Eqn. S31, showing
that our approximation is in good agreement with simulations. (Note that our approximation of
the density as linearly varying becomes less exact for slow motor speeds, leading to a discrepancy
between simulation results and our analytic approximation.) For comparison, we show the Langmuir
density ρ0 for the same parameter. Varying motor speed has no effect on ρ0 (right panel), but a
significant effect on ρc. To further illustrate the dependence of the center density on motor kinetic
parameters, we show in fig. 6 examples of motor density profiles from kMC simulations with varying
motor speed. Changing motor kinetic parameters can have a large effect on the center density in
the overlap, even for systems with identical binding kinetic parameters and ρ0.
Control of end accumulation by the total binding constraint
The length of the boundary layer at the overlap ends depends on motor binding and kinetic pa-
rameters as well as the overlap length (Eqn. S30). We compared the analytic approximation to
simulation results and found good agreement (fig. S10), allowing us to predict how varying experi-
mental control parameters will alter the degree of motor accumulation near the overlap ends (fig. 7).
We note that although the boundary layer lengths predicted for typical experimental parameters
are below the optical resolution limit of ≈ 250 nm, they could be distinguished by fits of fluorescent
motor intensity profiles to model predictions (c.f. figs. 1, 4, 6), similar to how fits to a Gaussian
fluorescence intensity profile allow sub-resolution localization of single molecules (41).
The total binding constraint means the the boundary layer length is determined as fraction of
the overlap length. In the single-filament model, the motor density profile in the MT minus-end
Motors on antiparallel MT overlaps
9
boundary layer is independent of MT length (17, 24), with important consequences: MT length
regulation by the kinesin-8 motor Kip3 (7, 16) depends on a motor density profile that maintains
the same functional form near the minus end as MT length changes. By contrast, on antiparallel
MT overlaps the density profile, boundary-layer length, and extent of motor accumulation at the
overlap ends change with overlap length (fig. 7). This means that a length regulation mechanism
identical to that of Kip3 cannot occur on antiparallel MT overlaps.
Conclusion
We studied a model of motor motion on antiparallel MT overlaps that incorporates motor binding
and unbinding, plus-end directed motor motion, and switching between filaments (fig. 1). Our
model is inspired by the experiments of Bieling, Telley, and Surrey on the motion of the kinesin-
4 motor Xklp1 on antiparallel MT overlaps (8). Our model is an extension of previous theory
that studied motor motion on a single MT with binding kinetics (5, 34), or motor motion on two
antiparallel filaments with lane switching, but no binding kinetics (35, 36). To our knowledge, this
is the first theoretical study of a two-lane TASEP model with oppositely oriented lanes, switching,
and binding/unbinding.
To compare to the BTS experiments, we used measured or estimated parameters (table 1).
Because PRC1 recruits motors directly into the overlap and motor binding to single MTs is much
weaker, we neglect motor binding to MTs outside of the overlap and set the minus-end flux param-
eter α = 0. Using the bulk motor unbinding rate koff as an upper bound on the plus-end unbinding
rate, we find β ≤ 2.7 × 10−3. This value is sufficiently small that it gives motor density profiles
indistinguishable from those with β = 0 (fig. S1). Therefore no-flux boundary conditions with
α = β = 0 are used to model the experiments. Simulated images of the motor distribution in an
overlap and simulated kymographs of motor trajectories are similar to those found experimentally
(fig. 1).
We derived analytical and approximate solutions of the continuum steady-state equations and
compared them to kMC simulation results (fig. S2). Focusing on the symmetric case where the
boundary conditions α and β are the same for both filaments in the overlap, we used both kMC
simulations and phase plane flows (fig. S3 -- S5) to determine the nonequilibrium phases possible
in our model (fig. 2). For low rate of motor switching between filaments in the overlap, we find the
low density, high density, low density-high density, and Meissner phases previously studied for the
single-lane case (5). In addition, for high switching rate we find a novel low density-high density-low
density-high density phase with three domain walls. We determined representative phase diagrams
for small and large switching rate (fig. 3).
We then studied the model in more detail both for the reference parameters and a high-
switching-rate parameter set (table S1). Results of kMC simulations (fig. 4, S6) agree well with
exact and approximate solutions of the continuum steady-state equations (fig. S2, S7).
In con-
trast to systems in which motors move on single filaments (7, 17 -- 22, 24), we find that antiparallel
overlaps with no-flux boundary conditions have total zero current. This leads to a total binding
constraint that the integral of the total motor density on a single filament must equal ρ0, the motor
density set by binding/unbinding equilibrium (fig. S8).
For the experimentally relevant low density-high density coexistence phase, the density profiles
are approximately piecewise linear (fig. S9). This motivates an analytic approximation to determine
the density profiles consistent with the total binding constraint and gives analytic expressions for
the overlap center motor density and the length of the boundary layer in which motors accumulate
near the overlap ends. We find that as a result of the total binding constraint, the motor density
Motors on antiparallel MT overlaps
10
at the center of the overlap is not determined solely by the motor binding equilibrium, but is also
controlled by the overlap length, motor speed, and filament switching rate (fig. 5, 6). These same
parameters control the length of the boundary layer at the overlap ends where motors accumulate
(fig. S10).
The mitotic spindle contains arrays of overlapping antiparallel MTs to which multiple motors
and crosslinkers bind (29). The surprising differences in motor density profiles between single
filaments and the antiparallel overlaps we study here are therefore of interest in the study of the
spindle midzone. For antiparallel overlaps, both the motor density far from the overlap ends and the
number of motors near the overlap ends can be tuned not just by motor binding kinetics but also by
motor speed, filament switching rate, and overlap length (fig. 7). The antiparallel filament geometry
gives biological systems additional handles to control motor density, its spatial distribution, and
therefore motor function. Motor density can affect recruitment of other proteins and MT dynamics.
Therefore, this previously undescribed mechanism of regulation of motor density along MTs may
offer advantages to the control of motor activity.
For single MTs, length regulation by the kinesin-8 motor Kip3 depends on a functional form
of motor density that is independent of MT length (7, 16). The physics we describe here means
that the motor density and extent of motor accumulation at the overlap ends depends on overlap
length. Therefore, length regulation of antiparallel MT overlaps must occur differently. In future
work, it will be of interest to understand how the motor density profiles on fixed-length overlaps
can be used to understand the length regulation of dynamic overlaps.
Acknowledgements
We thank Robert Blackwell, Matthew Glaser, and Loren Hough for useful discussions. This work
was supported by NSF grant DMR-0847685 and NIH grant GM110486 to MDB, fellowship to H-SK
provided by matching funds from the NIH/CU Biophysics Training Program, and facilities of the
Soft Materials Research Center under NSF MRSEC Grants DMR-0820579 and DMR-1420736.
References
1. Bray, D., 2000. Cell movements: from molecules to motility. Routledge.
2. Kolomeisky, A. B., 2015. Motor Proteins and Molecular Motors. CRC Press.
3. Chowdhury, D., 2013. Modeling Stochastic Kinetics of Molecular Machines at Multiple Levels:
From Molecules to Modules. Biophysical Journal 104:2331 -- 2341.
4. Helbing, D., 2001. Traffic and related self-driven many-particle systems. Rev. Mod. Phys.
73:1067 -- 1141.
5. Parmeggiani, A., T. Franosch, and E. Frey, 2004. Totally asymmetric simple exclusion process
with Langmuir kinetics. Phys. Rev. E 70:046101.
6. Leduc, C., K. Padberg-Gehle, V. Varga, D. Helbing, S. Diez, and J. Howard, 2012. Molecular
crowding creates traffic jams of kinesin motors on microtubules. PNAS 109:6100 -- 6105.
7. Varga, V., C. Leduc, V. Bormuth, S. Diez, and J. Howard, 2009. Kinesin-8 Motors Act Coop-
eratively to Mediate Length-Dependent Microtubule Depolymerization. Cell 138:1174 -- 1183.
Motors on antiparallel MT overlaps
11
8. Bieling, P., I. A. Telley, and T. Surrey, 2010. A Minimal Midzone Protein Module Controls
Formation and Length of Antiparallel Microtubule Overlaps. Cell 142:420 -- 432.
9. Goshima, G., R. Wollman, N. Stuurman, J. M. Scholey, and R. D. Vale, 2005. Length control
of the metaphase spindle. Curr. Biol. 15:1979 -- 1988.
10. Walczak, C. E., T. J. Mitchison, and A. Desai, 1996. XKCM1: A Xenopus Kinesin-Related
Protein That Regulates Microtubule Dynamics during Mitotic Spindle Assembly. Cell 84:37 --
47.
11. Rivero, F., B. Koppel, B. Peracino, S. Bozzaro, F. Siegert, C. J. Weijer, M. Schleicher, R. Al-
brecht, and A. A. Noegel, 1996. 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. J Cell Sci 109:2679 -- 2691.
12. Revenu, C., R. Athman, S. Robine, and D. Louvard, 2004. The co-workers of actin filaments:
from cell structures to signals. Nat. Rev. Mol. Cell Biol. 5:635 -- 646.
13. Dogterom, M., and S. Leibler, 1993. Physical aspects of the growth and regulation of micro-
tubule structures. Phys. Rev. Lett. 70:1347 -- 1350.
14. Drummond, D. R., 2011. Regulation of microtubule dynamics by kinesins. Seminars in Cell &
Developmental Biology 22:927 -- 934.
15. 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.
16. 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.
17. Hough, L. E., A. Schwabe, M. A. Glaser, J. R. McIntosh, and M. D. Betterton, 2009. Micro-
tubule depolymerization by the kinesin-8 motor Kip3p: a mathematical model. Biophys. J.
96:3050 -- 3064.
18. Reese, L., A. Melbinger, and E. Frey, 2011. Crowding of Molecular Motors Determines Micro-
tubule Depolymerization. Biophys. J. 101:2190 -- 2200.
19. Govindan, B. S., M. Gopalakrishnan, and D. Chowdhury, 2008. Length control of microtubules
by depolymerizing motor proteins. Europhys. Lett. 83:40006.
20. Johann, D., C. Erlenkamper, and K. Kruse, 2012. Length Regulation of Active Biopolymers
by Molecular Motors. Phys. Rev. Lett. 108:258103.
21. Melbinger, A., L. Reese, and E. Frey, 2012. Microtubule Length Regulation by Molecular
Motors. Phys. Rev. Lett. 108:258104.
22. Reese, L., A. Melbinger, and E. Frey, 2014. Molecular mechanisms for microtubule length
regulation by kinesin-8 and XMAP215 proteins. Interface Focus 4:20140031.
23. Tischer, C., P. R. ten Wolde, and M. Dogterom, 2010. Providing Positional Information with
Active Transport on Dynamic Microtubules. Biophys. J. 99:726 -- 735.
Motors on antiparallel MT overlaps
12
24. Kuan, H.-S., and M. D. Betterton, 2013. Biophysics of filament length regulation by molecular
motors. Phys. Biol. 10:036004.
25. Glunci´c, M., N. Maghelli, A. Krull, V. Krsti´c, D. Ramunno-Johnson, N. Pavin, and I. M. Toli´c,
2015. Kinesin-8 Motors Improve Nuclear Centering by Promoting Microtubule Catastrophe.
Phys. Rev. Lett. 114:078103.
26. Kurasawa, Y., W. C. Earnshaw, Y. Mochizuki, N. Dohmae, and K. Todokoro, 2004. Essential
roles of KIF4 and its binding partner PRC1 in organized central spindle midzone formation.
EMBO J. 23:3237 -- 3248.
27. Zhu, C., and W. Jiang, 2005. Cell cycle-dependent translocation of PRC1 on the spindle by
Kif4 is essential for midzone formation and cytokinesis. PNAS 102:343 -- 348.
28. Khmelinskii, A., C. Lawrence, J. Roostalu, and E. Schiebel, 2007. Cdc14-regulated midzone
assembly controls anaphase B. J Cell Biol 177:981 -- 993.
29. Fededa, J. P., and D. W. Gerlich, 2012. Molecular control of animal cell cytokinesis. Nat Cell
Biol 14:440 -- 447.
30. Subramanian, R., S.-C. Ti, L. Tan, S. A. Darst, and T. M. Kapoor, 2013. Marking and
Measuring Single Microtubules by PRC1 and Kinesin-4. Cell 154:377 -- 390.
31. Subramanian, R., E. M. Wilson-Kubalek, C. P. Arthur, M. J. Bick, E. A. Campbell, S. A. Darst,
R. A. Milligan, and T. M. Kapoor, 2010. Insights into Antiparallel Microtubule Crosslinking
by PRC1, a Conserved Nonmotor Microtubule Binding Protein. Cell 142:433 -- 443.
32. Bringmann, H., G. Skiniotis, A. Spilker, S. Kandels-Lewis, I. Vernos, and T. Surrey, 2004. A
Kinesin-like Motor Inhibits Microtubule Dynamic Instability. Science 303:1519 -- 1522.
33. Castoldi, M., and I. Vernos, 2006. Chromokinesin Xklp1 Contributes to the Regulation of
Microtubule Density and Organization during Spindle Assembly. Mol. Biol. Cell 17:1451 --
1460.
34. Parmeggiani, A., T. Franosch, and E. Frey, 2003. Phase Coexistence in Driven One-Dimensional
Transport. Phys. Rev. Lett. 90:086601.
35. Juh´asz, R., 2007. Weakly coupled, antiparallel, totally asymmetric simple exclusion processes.
Phys. Rev. E 76:021117.
36. Ashwin, P., C. Lin, and G. Steinberg, 2010. Queueing induced by bidirectional motor motion
near the end of a microtubule. Phys. Rev. E 82:051907.
37. Braun, M., Z. Lansky, G. Fink, F. Ruhnow, S. Diez, and M. E. Janson, 2011. Adaptive
braking by Ase1 prevents overlapping microtubules from sliding completely apart. Nat. Cell
Biol. 13:1259 -- 1264.
38. Lansky, Z., M. Braun, A. Ludecke, M. Schlierf, P. R. ten Wolde, M. E. Janson, and S. Diez,
2015. Diffusible Crosslinkers Generate Directed Forces in Microtubule Networks. Cell 160:1159 --
1168.
39. Bieling, P., I. Kronja, and T. Surrey, 2010. Microtubule Motility on Reconstituted Meiotic
Chromatin. Current Biology 20:763 -- 769.
Motors on antiparallel MT overlaps
13
40. 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.
41. Yildiz, A., J. N. Forkey, S. A. McKinney, T. Ha, Y. E. Goldman, and P. R. Selvin, 2003. Myosin
V Walks Hand-Over-Hand: Single Fluorophore Imaging with 1.5-nm Localization. Science
300:2061 -- 2065.
Motors on antiparallel MT overlaps
14
Figure Legends
Figure 1.
(A) Schematic of the model of motor motion on an antiparallel
Model and results overview.
microtubule overlap. Two filaments (green and blue) are modeled as 1D lattices with their plus
ends oppositely oriented. The filaments are labled R (L) if the plus end is pointing to the right
(left). Motors (red) bind to empty lattice sites with rate konc and unbind with rate koff . Bound
motors step toward the MT plus end with rate v (if the adjacent site toward the MT plus end is
empty) or switch to the other MT with rate s (if the corresponding site on the adjacent MT is
empty). At MT minus ends, motors are inserted at rate αv. At MT plus ends, motors are removed
at rate βv. (B) Simulated experimental images made from our kMC model. Green, motor density.
Red, overlap region. Scale bar, 5 µm. Simulations used to generate these images used the reference
parameter set (table 1) and the indicated bulk motor concentrations. (C) Simulated kymograph
made from our kMC model with motor spatial position on the horizontal axis and time increasing
down. Horizontal scale bar, 10 µm. Vertical scale bar, 5s. The simulations used to generate the
kymograph used the reference parameter set and 0.5 nM bulk motor concentration.
Figure 2.
Nonequilibrium phases. Left: motor density ρR(x) on MT with rightward-moving motors. Right:
total motor density ρR(x)+ρL(x) on both MTs in the overlap. The parameters used are the reference
parameter set of table 1 with a bulk motor concentration of 200 nM, except for the switching rate
which is 0.5 s−1 for the LHLH curve and 0.1 s−1 for all other curves and the boundary conditions
as noted below. LHLH: low density-high density-low density-high density coexistence (α = 0.3,
β = 0.9, orange). He: high-density end-accumulating phase (α = 0.95, β = 0.99, light blue). Hc:
high-density center-accumulating phase (α = 0.6, β = 0.95, dark blue). LHc:
low density-high
density center-accumulating coexistence (α = 0.3, β = 0.95, green). M: Meissner phase α = 0.6,
β = 0.4, yellow). Le:
low
density-high density end-accumulating coexistence (α = 0.1, β = 0.95, magenta). Lc: low-density
center-accumulating phase (α = 0.05, β = 0.1, red).
low-density end-accumulating phase (α = 0.3, β = 0.4, black). LHe:
Figure 3.
Phase diagrams. Left: low switching rate (0.1 s−1); right: high switching rate (0.5 s−1). Solid
lines indicate phase boundaries, and dotted lines boundaries between center-accumulating (v) and
end-accumulating (e) overlap motor density profiles. The bulk motor concentration is 200 nM, the
motor speed is 5 µm s−1, and other parameters are the reference values (table 1).
Figure 4.
Motor density profiles for the reference parameter set. Left: motor density ρR(x) on MT with
rightward-moving motors. Right: total motor density ρR(x) + ρL(x) on both MTs in the overlap.
Parameters are the reference parameter set of table 1 with the bulk motor concentrations indicated
in the legend.
Figure 5.
Motor density at the center of the overlap. Left: variation with bulk motor concentration; right:
variation with motor speed. Points indicate simulation results and dashed lines theoretical pre-
Motors on antiparallel MT overlaps
15
diction from Eqn. S31. The red line shows the value of the Langmuir density ρ0 for the same
parameters.
Figure 6.
Dependence of density profiles on motor speed. Left: motor density ρR(x) on MT with rightward-
moving motors. Right: total motor density ρR(x) + ρL(x) on both MTs in the overlap. Varying
motor speed can significantly alter the motor density at the center of the overlap. These simulations
use the large-S parameter set of table S1 with bulk motor concentration c = 200 nM and the motor
speeds indicated in the legend.
Figure 7.
Control of boundary layer length. Left: boundary layer length as a fraction of total overlap length,
shown as a function of bulk motor concentration and motor speed. Right: boundary layer length as
a function of overlap length for the reference parameter set and varying bulk motor concentration.
The boundary layer length is the distance at each end of the overlap where significant motor
accumulation occurs and is determined by Eqn. S30.
Motors on antiparallel MT overlaps
16
Figure 1: Model and results overview. (A) Schematic of the model of motor motion on an antipar-
allel microtubule overlap. Two filaments (green and blue) are modeled as 1D lattices with their
plus ends oppositely oriented. The filaments are labled R (L) if the plus end is pointing to the right
(left). Motors (red) bind to empty lattice sites with rate konc and unbind with rate koff . Bound
motors step toward the MT plus end with rate v (if the adjacent site toward the MT plus end is
empty) or switch to the other MT with rate s (if the corresponding site on the adjacent MT is
empty). At MT minus ends, motors are inserted at rate αv. At MT plus ends, motors are removed
at rate βv. (B) Simulated experimental images made from our kMC model. Green, motor density.
Red, overlap region. Scale bar, 5 µm. Simulations used to generate these images used the reference
parameter set (table 1) and the indicated bulk motor concentrations. (C) Simulated kymograph
made from our kMC model with motor spatial position on the horizontal axis and time increasing
down. Horizontal scale bar, 10 µm. Vertical scale bar, 5s. The simulations used to generate the
kymograph used the reference parameter set and 0.5 nM bulk motor concentration.
Motors on antiparallel MT overlaps
17
Figure 2: Nonequilibrium phases. Left: motor density ρR(x) on MT with rightward-moving motors.
Right: total motor density ρR(x) + ρL(x) on both MTs in the overlap. The parameters used are
the reference parameter set of table 1 with a bulk motor concentration of 200 nM, except for the
switching rate which is 0.5 s−1 for the LHLH curve and 0.1 s−1 for all other curves and the
boundary conditions as noted below. LHLH: low density-high density-low density-high density
coexistence (α = 0.3, β = 0.9, orange). He: high-density end-accumulating phase (α = 0.95,
β = 0.99, light blue). Hc: high-density center-accumulating phase (α = 0.6, β = 0.95, dark blue).
low density-high density center-accumulating coexistence (α = 0.3, β = 0.95, green). M:
LHc:
Meissner phase α = 0.6, β = 0.4, yellow). Le:
low-density end-accumulating phase (α = 0.3,
β = 0.4, black). LHe: low density-high density end-accumulating coexistence (α = 0.1, β = 0.95,
magenta). Lc: low-density center-accumulating phase (α = 0.05, β = 0.1, red).
Figure 3: Phase diagrams. Left: low switching rate (0.1 s−1); right: high switching rate (0.5 s−1).
Solid lines indicate phase boundaries, and dotted lines boundaries between center-accumulating (v)
and end-accumulating (e) overlap motor density profiles. The bulk motor concentration is 200 nM,
the motor speed is 5 µm s−1, and other parameters are the reference values (table 1).
Motors on antiparallel MT overlaps
18
Figure 4: Motor density profiles for the reference parameter set. Left: motor density ρR(x) on
MT with rightward-moving motors. Right: total motor density ρR(x) + ρL(x) on both MTs in the
overlap. Parameters are the reference parameter set of table 1 with the bulk motor concentrations
indicated in the legend.
Figure 5: Motor density at the center of the overlap. Left: variation with bulk motor concentration;
right: variation with motor speed. Points indicate simulation results and dashed lines theoretical
prediction from Eqn. S31. The red line shows the value of the Langmuir density ρ0 for the same
parameters.
Motors on antiparallel MT overlaps
19
Figure 6: Dependence of density profiles on motor speed. Left: motor density ρR(x) on MT with
rightward-moving motors. Right: total motor density ρR(x) + ρL(x) on both MTs in the overlap.
Varying motor speed can significantly alter the motor density at the center of the overlap. These
simulations use the large-S parameter set of table S1 with bulk motor concentration c = 200 nM
and the motor speeds indicated in the legend.
Figure 7: Control of boundary layer length. Left: boundary layer length as a fraction of total
overlap length, shown as a function of bulk motor concentration and motor speed. Right: boundary
layer length as a function of overlap length for the reference parameter set and varying bulk motor
concentration. The boundary layer length is the distance at each end of the overlap where significant
motor accumulation occurs and is determined by Eqn. S30.
|
1005.4861 | 1 | 1005 | 2010-05-26T16:11:11 | Wave Propagation in Lipid Monolayers | [
"physics.bio-ph",
"cond-mat.soft"
] | Sound waves are excited on lipid monolayers using a set of planar electrodes aligned in parallel with the excitable medium. By measuring the frequency dependent change in the lateral pressure we are able to extract the sound velocity for the entire monolayer phase diagram. We demonstrate that this velocity can also be directly derived from the lipid monolayer compressibility and consequently displays a minimum in the phase transition regime. This minimum decreases from v0=170m/s for one component lipid monolayers down to vm=50m/s for lipid mixtures. No significant attenuation can be detected confirming an adiabatic phenomenon. Finally our data propose a relative lateral density oscillation of \Delta\rho/\rho ~ 2% implying a change in all area dependent physical properties. Order of magnitude estimates from static couplings therefore predict propagating changes in surface potential of 1-50mV, 1 unit in pH (electrochemical potential) and 0.01{\deg}K in temperature and fall within the same order of magnitude as physical changes measured during nerve pulse propagation. These results therefore strongly support the idea of propagating adiabatic sound waves along nerves as first thoroughly described by Kaufmann in 1989 and recently by Heimburg and Jackson, but claimed by Wilke already in 1912. | physics.bio-ph | physics | Wave Propagation in Lipid Monolayers
J. Griesbauer1, A. Wixforth1, M.F. Schneider1*
1 University of Augsburg, Experimental Physics I, D-86159 Augsburg, Germany
* Corresponding author:
Matthias F. Schneider
Current:
Phone: +49-821-5983311, Fax: +49-821-5983227
[email protected]
University of Augsburg, Experimental Physics I,
Biological Physics Group
Universitätstr. 1
D-86159 Augsburg, Germany
New Address (September 1, 2009):
[email protected]
Boston University, Dept. of Mechanical Engineering
110 Cummington St
Boston-Massachusetts, USA
Abstract
Sound waves are excited on lipid monolayers using a set of planar electrodes aligned in
parallel with the excitable medium. By measuring the frequency dependent change in the
lateral pressure we are able to extract the sound velocity for the entire monolayer phase
diagram. We demonstrate that this velocity can also be directly derived from the lipid
monolayer compressibility and consequently displays a minimum in the phase transition
regime. This minimum decreases from v0=170m/s for one component lipid monolayers down
to vm=50m/s for lipid mixtures. No significant attenuation can be detected confirming an
adiabatic phenomenon. Finally our data propose a relative lateral density oscillation of Δρ/ρ ~
2% implying a change in all area dependent physical properties. Order of magnitude estimates
from static couplings therefore predict propagating changes in surface potential of 1-50mV, 1
unit in pH (electrochemical potential) and 0.01°K in temperature and fall within the same
order of magnitude as physical changes measured during nerve pulse propagation. These
results therefore strongly support the idea of propagating adiabatic sound waves along nerves
1
as first thoroughly described by Kaufmann in 1989 and recently by Heimburg and Jackson,
but claimed by Wilke already in 1912.
2
Introduction
The lipid monolayer is ubiquitously present in biology as one half of the cell, organelle or
vesicle membrane. A complete understanding of the physical properties of lipid monolayers is
therefore of fundamental interest in order to understand its role for biological bilayer systems.
Studying sound propagation in insoluble organic films the lipid monolayer obtains its
attraction from the fact that most physical properties (lateral pressure, area per molecule,
compressibility, surface potential, temperature etc.) are easily accessible (1). Consequently
lipid monolayers have been investigated very intensively from various viewpoints of physics
and physical chemistry (see (1-4) and references therein). On the other hand only very little
studies addressed internal excitations or wave propagation within lipid monolayers (5). If
however two dimensional adiabatic excitations are present in simple lipid monolayer systems
their absence in complex biological lipid membranes appears very unlikely.
Here, we present both experimental and theoretical studies on adiabatic sound wave
propagation in lipid monolayers. Experimentally, lipid monolayers are excited by an
alternating in-plane electric field (in-plane excitation electrodes, IPE) originating from a set of
laterally patterned electrodes (interdigital transducers, IDT). For certain distinct frequencies
f , we find pronounced changes in the lateral pressure which suggests electrical excitation with
an underlying resonance phenomenon. This in turn allows to extract in detail the variable
sound velocity of the propagating wave by matching the excitation frequency f with the
distance d between two planar electrodes, defining an imposed wave length.
Thermodynamic analysis of the problem reveals that both the lateral pressure change as well
as the sound velocity for the different thermodynamic states can be directly derived from the
lipid monolayer phase diagram and gives clear evidence for the existence of a propagating,
adiabatic sound wave. Order of magnitude estimates suggest that the origin of the decoupled
propagation may be a consequence of the different sound velocities in 2D and 3D as well as
the increased heat conductivity of the boundary water layer.
Materials and Methods
Lipids
1,2-
and
(DPPG)
1,2-Dipalmitoyl-sn-Glycero-3-[Phospho-rac-(1-glycerol)]
Dipalmitoyl-sn-Glycero-3-Phosphocholine (DPPC) dissolved in Chloroform were purchased
from Avanti Polar Lipids (Birmingham, Al. USA) and used without further purification.
All measurements were done on a standard film balance with a heat bath (NIMA, Coventry,
England) modified by a small stage to position the excitation chips with the IDT structure.
Standard isotherms could be recorded and the regulation circuit of the film balance allowed it
to hold a specific pressure of the monolayer for measuring the area expansion coefficient
α (T).
The IPE chips were produced by standard lithography method, using LiTaO3 as substrate for
the gold electrodes. The interdigitated gold electrodes (finger spacing w ≈ 10µm) were
connected to an RF amplifier (ZHL-2010+, MINI-CIRCUITS, Brooklyn, NY, USA) which
was driven by a signal generator (SML 01, Rhode&Schwarz, Munich, Germany). Such chips
are routinely used in our lab for surface acoustic wave studies, including microfluidic and
sensor applications. Here, however, we operate the chips to generate RF electric fields
coupling to the lipid monolayers only and at frequencies where no surface waves of the
substrate are resonantly excited.
To obtain lipid mixtures, lipids dissolved in chloroform were mixed in the desired
proportions. Before spreading the lipid monolayers on the air water interface, the chip and the
Wilhelmy plate were placed to the surface of the trough as described in Fig. 2. Frequency
sweeps between f=0 and f=27 MHz were performed during 30 minutes, whereas the pressure
was recorded simultaneously.
3
Results and Discussion
Theory
Fundamental Thermodynamic Relations
In Einstein’s first publication in 1901 he approved our crucial assumption: applying a Carnot
cycle to a free surface he demonstrated that the interface of a water drop must have its own
heat (heat capacity) and therefore its own entropy SI (6, 7) . Our experiments as well as order
of magnitude estimates will demonstrate that a lipid monolayer at the air/water interface has
its own entropy SI as well and has therefore to be considered as an independent
thermodynamic system. To realize the meaning of SI for the experiment it is very helpful to
recall that it is the second derivative of SI with respect to a thermodynamic variable x that is
related to the generalized susceptibilities by (8)
k
B
⎛
⎜⎜
⎝
I
S
2
∂
2
x
∂
i
−1
⎞
−≡⎟⎟
⎠
k
B
⎛
⎜⎜
⎝
x
∂
i
X
∂
i
⎞
⎟⎟
⎠
.
iJX
≠
(1)
Here kB denotes the Boltzmann constant and
S
μπ
∂
,...
,
TT
x
∂
i
corresponding thermodynamic force. For xi = A and constant temperature, this results in
(e.g. Xi =
−=
X
I
i
) the
where
κ
T
−=
A
1
∂
A π
∂
T
k
B
I
S
2
∂
A
2
∂
⎛
⎜⎜
⎝
−1
⎞
−=⎟⎟
⎠
T
TAk
κ
T
B
,
(2)
(3).
This is the lateral isothermal compressibility of the lipid monolayer, which can be directly
derived from the pressure-area isotherm (Figure 1). Hence, the mechanical properties of the
lipid monolayer represent a measure of the susceptibility, i.e. inverse curvature of the entropy
potential. A maximum in κT as observed in the phase transition regime corresponds to a “flat”
entropy potential. It should be noted, that different boundary conditions (keeping T, q, μ or N,
etc. constant), will result in different potentials like different 1D projections of the same n-
dimensional potential.
Estimation of
In principle, the velocity c0 could be estimated within an order of magnitude by using κT. The
easy accessibility of the thermal expansion coefficient α however, enables us to get a better
ISκ using the thermodynamic relation (9)
estimate of the adiabatic compressibility
2
T
A
∂
⎛
⎞
⎜
⎟
Ac
T
∂
⎝
⎠
π
π
κκ
≈
T
S I
from
Tκ
ISκ
(4)
−
,
4
B
where cπ represents the heat capacity at constant pressure and
A
1
∂
A ∂
T
π
expansion coefficient. Even though this relationship is used quite commonly it should be
noted that it strictly holds only for a system defined by π, A and T and will carry additional
terms, when charge, dipole, chemical potential, etc. are included. Nevertheless, extracting κT
and απ from the experiment and approximating cπ by using the experimentally established
correspondence between change in enthalpy ΔH and area ΔA (9)
the isobaric
α
=
(5a)
and therefore
,
(5b)
A
H
γ
Δ≈Δ
⎛≈Δ
c
π γ
⎜
⎝
dA
dT
⎞
⎟
⎠
π
Equation 4 can now be used to estimate the propagation velocity c0 of a sound wave, which in
the linear case is given from the fundamental thermodynamic relation
.
c
0 =
ρκ/1
IS
(6)
Tκ
Here, ρ is the area per molecule, an experimentally well controllable quantity. Since both
as well as
Tκ exhibit a maximum in the phase transition regime the sound velocity is expected
to undergo a minimum near the isothermal phase transition of the lipid monolayer. Clearly,
inseparable from the excited sound wave a temperature wave must propagate along the lipid
monolayer as well. Knowing the propagating area density oscillation an estimate of the
accompanied temperature change can be extracted from the isothermal expansion and the heat
capacity of the monolayer by (10)
T
=Δ
Tc
0
vA
ρ
≈Δ
vTc
,
−1
ργ
Δ
0
(7)
α
π
c
π
where v is the particle velocity and Eq. 5b has been applied as well. Again, even commonly
used, this strictly holds only for a system defined by π, A and T.
Experiments
Excitation of Sound Waves by Planar Electrodes
When a thermodynamic system is forced out of its equilibrium position, the conservation of
entropy requires a propagation phenomenon. In this sense the observation of sound
propagation in lipid monolayers would provide additional support that the lipid membrane
interface can be considered as a “closed” two dimensional system, fairly well decoupled from
the surrounding bulk bath. Qualitatively a flat thermodynamic potential implies weak
restoring forces and therefore a lower propagation velocity as for steep potentials.
In our experiments the excitation of propagating sound waves along the lipid membrane was
accomplished by incorporating a chip with a planar array of gold electrodes (in-plane
excitation, IPE) in the plane of a negatively charged DPPG monolayer (Figure 2).
Traditionally, such IDTs are used as filters in RF applications or to create acoustic streaming
in microfluidic systems (11, 12). Another application of IDT are sensors comparable to the
5
well known quartz crystal microbalance (QCM)(13, 14). Here, however, we use such
electrodes to excite in-plane waves on soft interfaces at various frequencies by electro-
mechanical coupling to a polarisable membrane. We employ the fact that a lateral density
creates a net pressure increase Δπ . This is due to a nonlinear
oscillation
B ω
t
cos
ρρ
=
⋅+
0
pressure-area relationship π(A) and can be calculated with the help of the isotherm by simply
integrating the pressure change over one period:
π
=Δ
ω
2
π
∫
(
π
0
ω
2
π
tA
))(
dt
−
(
π
A
)
The same expression can be written in terms of the mass density ρ :
(8a).
(8b).
=Δ
π
ω
2
π
ω
2
π
∫
(
π
0
ρ
0
m
Lipids
B
t
cos
⋅+
ω
)
dt
−
(
π
m
Lipids
ρ
0
)
Here, ω is the angular frequency of the travelling wave, B its amplitude,
0ρ the lateral
Lipidsm
density in rest and
the mass of the lipids forming the monolayer. The first term on the
right side of Eq. 8 represents the time averaged pressure change due to the modulation in the
interface, while the second term denotes the undisturbed lipid monolayer.
If adiabatic propagation would indeed take place, Eq. 8 predicts a net increase in lateral
pressure. The pressure spectrum of a DPPG monolayer shown in Fig. 3 confirms that such an
increase takes place indeed. The average change in lateral pressure πΔ detected during the
excitation of the wave is plotted as a function of the stimulating frequency applied to the chip
for three different configurations. Between f = 100 kHz and f = 27 MHz a pure water surface
does not produce any significant response. Nevertheless at a distance of 2cm at around f=33
MHz pure water may also show a response. The origin of this signal is presently still
unknown but may be attributed to surface capillary water waves (15). The presence of a
DPPG monolayer on this surface, however, results in a pronounced variations in πΔ around
f = 11 MHz. Importantly the response at f = 11MHz is also not visible on the pure water
surface at shorter distances between excitation and detection. We therefore conclude, that the
change of πΔ around f = 11MHz can ubiquitously be attributed to the presence of the lipid
membrane and will be used for further data interpretation.
Attenuation and Amplitude of the Wave
Considering the macroscopic (15cm) distance between excitation and detection, the
propagation of the wave does not seem to be significantly attenuated. To experimentally
verify this finding, we measured the change in lateral pressure πΔ as a function of distance
from the source of excitation (see inset of Fig. 3). It turns out that the experimental data can
xπ
bx
(
)(
2)
be well fitted by a polynomial
, but not with an exponential decay
−−−∝
Δ
function. This supports the idea of an adiabatic wave with decay due only to the geometry of
the system but little dissipation.
Finally, the amplitude of the excited wave can be estimated by comparing the experimentally
observed πΔ to Eq. 8 and reveals a density variation B ≈ 0.02ρ0 (± 0.005ρ0). This in turn
6
would result in a density modulation amplitude of about B ~0,3 ·10-7kg/m2, which is well
within reasons for lipid monolayers.
Sound velocity from lateral pressure changes
The observed πΔ clearly exhibits a frequency dependence (Fig. 3) with an apparent
resonance like feature at f = 11MHz, indicating significant excitation at this point. It appears,
that sound velocity c0, electrode spacing d and stimulating frequency ν0 are ideally matched at
exactly this frequency in order to provide effective excitation. Taking the finger spacing
between two electrodes d = 12 µm (therefore λ = 24 µm) and the resonance frequency ν0 =
11MHz from the experiment, we find a propagation velocity of c0=λν0 = 260m/s.
Following the same procedure, we repeated the experiment along the entire isotherm also
including the phase transition regime and extracted the corresponding sound velocity. It turns
out that the different thermodynamic states of the monolayer indeed exhibit different
excitation frequencies ν0 as can be seen in Fig. 4 for three different surface pressures.
A critical test whether the origin of the observed pressure change is indeed a lateral density
oscillation arises from the fact that Eq. 8 predicts a negative change in πΔ in and below the
maximal compressibility, while πΔ is positive for pressures above the maximum. Clearly,
this behaviour is qualitatively and quantitatively reproduced in Figure 4 and does therefore
strongly support our assumption of a propagating sound wave along the interface.
Sound velocity from the monolayer compressibility
The experiments described above enabled us to extract the sound velocities along the entire
isotherm. If this represents indeed a two dimensional sound wave the velocity c0 should
directly depend on the adiabatic lateral compressibility (Eq. 6). Even though κT will produce
the right order of magnitude for the sound velocity, we exploit the fact that all other
susceptibilities appearing in Eq. (4) can be extracted from the monolayer isotherm as well and
will provide the basis to calculate a more accurate κS. Therefore απ was measured from the
A(T) isobars for different lateral pressures, to calculate a more accurate approximation of κS.
cπ was calculated by applying Eq. 5, which should give a good estimate at least close to the
phase transition regime (9). Finally, comparing the propagation velocities as being calculated
from Eq. 6 and the measured sound velocities using c0=λν0 , we find both qualitative and
quantitative agreement within 10% or less (Fig. 5a) confirming the existence of a propagating
sound wave.
In particular, the minimum in propagation speed as a consequence of the maximum in κT is
resolved. The same qualitative behaviour is observed for a mixture of DPPG/DPPC (1:10).
Again excellent quantitative agreement is achieved between c0 calculated from Eq.4 and the
experimental pressure spectra (Fig. 5b). For this lipid, however, the minimum velocity
predicted from the isotherm and confirmed experimentally is only c0 = 50m/s corresponding
to the phase transition regime between liquid expanded and liquid condensed phase. This
finding illustrates that both physical parameters (lateral pressure, temperature etc.) and the
composition of the monolayer control the propagation velocity.
7
On the coupling between monolayer and bulk
The only reasonable propagation mechanism for our system is one in which heat and entropy
SI (Eq. 1-3) of the interface are approximately conserved (no exponential decay in wave
energy). Therefore, the fact that only weakly damped wave propagation is experimentally
observed (see inlet of Fig. 3) calls for a decoupling between monolayer and bulk. Although,
this is an experimental result and strongly supported by Einstein’s early work on the heat of
surfaces (7), we would like to outline another argument that assumes coupling but will
demonstrate its insignificance at the same time. Following closely the book of Landau &
Lifshitz (Vol 6) chapter II, V and VIII (10), we present an order of magnitude estimate of why
the lipid monolayer may be decoupled from the bulk.
When a wave propagates in media 1 (monolayer) at the interface to an adjacent media 2 (bulk
water), the reflection coefficient R depends on the angle of incident θ, the density ρ and sound
velocity c of the two media (10, VIII).
2
(9)
R
=
⎡
⎢
⎢
⎣
c
ρ
22
c
ρ
22
cos
−
ρθ
1
cos
+
ρθ
1
c
(
2
1
c
(
2
1
−
−
c
2
2
c
2
2
sin
sin
2
2
)
θ
)
θ
⎤
⎥
⎥
⎦
If the incident wave forms an angle of less than the critical angle θc, where sinθc = c1/c2 (10,
VIII), the entire wave is reflected (total internal reflection). Taking c1 ~ 100 m/s and c2 ~1500
m/s we obtain a critical angle of θ0 ≈ 5° for our arrangement. Lateral waves excited within the
lipid monolayer will therefore be completely reflected. In this situation Landau (10, II)
demonstrates that oscillations parallel to the interface of exponential decay in amplitude are
created. These oscillations, however, do not provide significant dissipation as their
exponential penetration into the bulk only resembles the energy distribution of the oscillating,
propagating sound wave around the interface and must not be mistaken with the actual
transport of energy out of the system, but as inseparable from the wave.
Similar arguments hold for the dissipation of heat. If the heat remains within the typical
spatial extension of the sound wave, dissipation cannot be significant. As the temperature
variations in the system are a direct consequence of the wave oscillations the temperature
changes perpendicular to the interface must be exponential as well (10, V)
exp(
−
z
TT
=
0
⎡
⎤
ω
ω
t
zi
)
exp(
ω
⎢
⎥
2
2
χ
χ
⎣
⎦
where ω = 2πf, and χ is the thermometric conductivity which can be calculated from the heat
capacity cP, the density ρ and the thermal conductivity k of the system (10, V)
(10a)
,
−
)
χ
=
k
Pc
⋅
.
ρ
(10b)
Following Landau (10, II) the extension or viscous penetration depth δ in z-direction (Fig. 6)
can be estimated from
δ
=
2
η
ρω
,
(11)
where η is the dynamic and η/ρ the kinematic viscosity of water. Using standard numbers for
water, this depth turns out to be δ ≈ 300nm. In order to estimate whether thermal diffusion
8
can add significantly to the dissipation process we need to compare δ with the thermal
penetration length ξ
2
χ
ω
ξ
=
,
(12)
,
(13)
where χ denotes the thermometric conductivity (see Eq. 10b). This length scale describes over
which distance a significant change in temperature takes place. Using cp ≈ 4 kJ(kgK)-1 and k ≈
0.6 J(mKs)-1 we arrive at ξ ≈ 100nm. Since δ ≥ ξ the heat is, even when assuming coupling
to the bulk, unable to “escape” the spatial extension of the sound wave within the timescale of
compression (0.5·10-7s here). In other words Eq. 10a simply describes the reversible
temperature oscillations inside the sound wave.
Finally, heat could also dissipate within the monolayer plane. This is to be predicted from a
two dimensional shear viscosity observed in lipid monolayers. To estimate the corresponding
loss inside the monolayer, we may also compare the length scale l of heat dissipation within
−1 to the wavelength λ. Only if the heat
the time of one compression/expansion cycle t = f0
expands (diffuses) during t over a distance l, which is of the same order of magnitude as the
wavelength λ, significant dissipation is to be expected. According to Landau, l can be
calculated from the solution of the general equation of heat transfer (10, V):
χτ≈2l
where χ is the aforementioned thermometric conductivity (Eq. 10b). Unfortunately, to our
knowledge, no numbers for the heat conductivity kI of lipid monolayers exist. However, kI
may be approximated by the heat conductivity of interfacial or boundary water, which has
experimentally been observed on lipid membranes (16, 17). Using kI ≈ 6 J/mKs (which is one
order of magnitude larger than for bulk water) (18), cp ≈ 10 kJ/kgK (19), and τ ≈ t the lateral
extension of heat due to thermal diffusion l ≈ 200nm, which is ~ 100 times smaller than the
wavelength (λ ~ 24μm). Therefore, the heat cannot “escape” the propagating wave in the
lateral direction either.
To summarize this paragraph, we would like to state that neither mechanical nor thermal
coupling properties support the admittedly intuitive prejudice of a strongly attenuated (since
coupled) wave, but are in agreement with our experimental observation of a propagating, only
weakly attenuated sound wave.
Conclusion and Biological Impact
A new approach to excite and detect acoustic waves in lipid monolayers is presented.
Moreover, the existence of adiabatic sound waves is experimentally confirmed and the
corresponding sound velocities are extracted from our measurements. Comparison of our
findings with theoretical predictions provide excellent agreement and reveal velocities
between 300 m/s and 170 m/s for one component, and 300 m/s to 50 m/s for two component
systems, with a distinct minimum in the phase transition regime. Considering the simplicity of
our system, these values are in very good agreement with reported propagation velocities of
action potentials in nerves, ranging between 10 – 100 m/s and depending not only on
myelination but also on temperature, thickness, sodium concentration etc. (20-23).
Our results provide an explanation of the well known, yet still unresolved problem of a
temperature variation which is observed during action potentials (24-27). Our findings even
predict that such reversible changes must occur. Quantitative estimates in ΔΤ(t) calculated
9
from static experiments and the modulation in area density ρ(t) using Eq. 7, propose
temperature variations in the lipid monolayer of 0.01°K. Propagating changes in surface
potential are expected as well. Taking surface potential measurements from static, isothermal
experiments (2), the observed variation in area density ρ(t) predicts a change in ΔU between 1
mV and 50 mV propagating along the surface of the lipid monolayer. Our results are therefore
in support of the idea of propagating sound waves in biological membranes as first discussed
by Wilke (28) and Wilke and Atzler (29) in 1912, first thoroughly described by Kaufmann in
1989 (30, 31), and recently discussed by Heimburg and Jackson in 2005 (24). The reported
changes in temperature and pressure observed during nerve-pulse propagation (24-27) are at
least in qualitative agreement with our predictions.
Acknowledgement
We thank Dr. K. Kaufmann (Göttingen) for very helpful discussions and highly recommend
the reader to consult his earlier work [s Ref. 30 and 31]. MFS likes to personally thank K.
Kaufmann who inspired him to work in this field and introduced him to the thermodynamic
origin of propagation along membranes and nerves.
Financial support by the Bundesministerium für Bildung und Forschung is gratefully
acknowledged. MFS likes to thank the Bavarian Science Foundation for financial support.
This work has also been partially funded by the German Excellence Initative ‘NIM’.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
Gaines, G. L. 1966. Insoluble Monolayers at Lipid-Gas Interfaces. Interscience,
New York.
Möhwald, H. 1995. Structure and Dynamics of Membranes. Elsevier: Amsterdam.
Albrecht, O., H. Gruler, and E. Sackmann. 1978. Polymorphism of phospholipid
monolayers. Journal de Physique I 39:301-313.
McConnell, H. M. 1991. Structures and Transitions in Lipid Monolayer at the
Air-Water Interface. Annu Rev Phys Chem 42:171 - 195.
Frey, W., and E. Sackmann. 1992. Solitary waves in asymmetric soap films.
Langmuir 8:3150-3154.
Einstein, A. 1901. Folgerungen aus den Capillaritätserscheinungen. Ann d Phys
4:513-523.
Einstein, A. 1910. Theorie der Opaleszenz von homogenen Flüssigkeiten und
Flüssigkeitsgemischen in der Nähe des kritischen Zustandes. Ann d Phys 25:205-
226.
Landau, L. D., and E. M. Lifschitz. 1987. Course of Theoretical Physics Vol. 5.
Butterworth-Heinemann.
Heimburg, T. 1998. Mechanical aspects of membrane thermodynamics.
Estimation of the mechanical properties of lipid membranes close to the chain
melting transition from calorimetry. Biochim Biophys Acta 1415:147-162.
Landau, L. D., and E. M. Lifschitz. 1987. Course of Theoretical Physics Vol. 6.
Butterworth-Heinemann.
Schneider, M. F., S. W. Schneider, V. M. Myles, U. Pamucki, Z. Guttenberg, K.
Sritharan, and A. Wixforth. 2007. An Acoustically Driven Microliter Flow
Chamber on a Chip (µFCC) for Cell/Cell and Cell/Surface Interaction Studies.
ChemPhysChem. 9:651-655.
10
13.
14.
18.
Schneider, S. W., S. Nuschele, A. Wixforth, C. Gorzelanny, A. Alexander-Katz,
R. R. Netz, and M. F. Schneider. 2007. Shear-induced unfolding triggers adhesion
of von Willebrand factor fibers. Proc Natl Acad Sci U S A 104:7899-7903.
Janshoff, A., H. J. Galla, and C. Steinem. 2000. Piezoelectric Mass-Sensing
Devices as Biosensors-An Alternative to Optical Biosensors? Angew Chem Int Ed
Engl 39:4004-4032.
Freudenberg, J., M. von Schickfus, and S. Hunklinger. 2001. A SAW
immunosensor for operation in liquid using a SiO/sub 2/ protective layer.
Elsevier. Sensors & Actuators B Chemical:1-3.
Il'ichev, A. 1998. Self-Channeling of surface water waves in the presence of an
additional surface pressure. Eur J Mech B 18:501-510
16. K. Åman, E. Lindahl, O. Edholm, P. Håkansson, P. Westlund. 2003. Structure
and Dynamics of Interfacial Water in an Lα Phase Lipid Bilayer from Molecular
Dynamics Simulations. Biophysical Journal, Vol. 84, 102-115.
17. V.Luzzati, F. Husson. 1962. The Structure Of The Liquid-Crystalline Phases Of
Lipid-Water Systems. J. Cell. Biol., Vol. 12, 207-219.
J. S. Clegg, W. Drost-Hansen. 1991. On the biochemistry and cell physiology of
water. P. W. Hochachka and T. P. Mommsen, 1–23.
19. A. Blume. 1983. Apparent Molar Heat Capacities of Phospholipids in Aqueous
Dispersion. Effects of Chain Length and Head Group Structure. Biochem., Vol.
22 (23), 5436–5442.
20. A. F. Huxley, R. Stampfli. 1949. Evidence for saltatory conduction in peripheral
myelinated nerve fibres. J. Physiol., Vol. 108, 315-339.
21. H. P. Ludin, F. Beyeler. 1977. Temperature dependence of normal sensory nerve
action potentials. J. Neurol., Vol. 216(3), 173–180.
J. B. Hursh. 1939. Conduction Velocity and Diameter of Nerve Fibres. Am. J.
Physiol., Vol. 127, 131-139.
23. W. L. Hardy. 1973. Propagation Speed in Myelinated Nerve. Biophys. J., Vol. 13,
1054-1070.
24. Heimburg, T., and A. D. Jackson. 2005. On soliton propagation in biomembranes
and nerves. Proc Natl Acad Sci U S A. 102:9790-9795.
Ritchie, J. M., and R. D. Keynes. 1985. The production and absorption of heat
associated with electrical activity in nerve and electric organ. Q Rev Biophys
18:451-476.
Tasaki, I. 1995. Mechanical and thermal changes in the Torpedo electric organ
associated with its postsynaptic potentials. Biochem Biophys Res Commun
215:654-658.
27. Kim, G. H., P. Kosterin, A. L. Obaid, and B. M. Salzberg. 2007. A Mechanical
Spike Accompanies the Action Potential in Mammalian Nerve Terminals.
Biophys J.
28. Wilke, E. 1912. Das Problem der Reizleitung im Nerven vom Standpunkt der
Wellenlehre aus betrachtet. Pflügers Arch 144:35-38.
29. Wilke, E., and E. Atzler. 1912. Experimentelle Beiträtge zum Problem tier
Reizleitung im Nervem. Pflügers Arch 146:430-446.
30. Kaufmann, K. 1989. On the role of the phospholipid membrane in free energy
coupling. Caruaru, Brazil (http://membranes.nbi.dk/ Kaufmann/pdf/ Kaufmann
book5 ed.pdf).
31. Kaufmann, K. 1989. Action Potentials and Electrochemical Coupling in the
Macroscopic Chiral Phospholipid Membrane. Caruara, Brazil
(http://membranes.nbi.dk/ Kaufmann/pdf/ Kaufmann book4 ed.pdf).
22.
25.
26.
12.
15.
11
Figure Captions
Fig. 1 Isotherms of lipid monolayers for 1,2-Dipalmitoyl-sn-Glycero-3-[Phospho-rac-(1-
for a mixture of 1,2-Dipalmitoyl-sn-Glycero-3-
glycerol)]
(DPPG) at 14°C and
Phosphocholine (DPPC) to DPPG as 10:1 at 20°C. Both Isotherms show maxima in the
compressibility at about 8 mN/m for DPPG and at about 2,5 mN/m for the mixture, whereas
the later clearly corresponds to the phase transition regime. The inset shows the resulting
compressibility of the corresponding monolayer. As for the in-plane excitation (IPE) setup, a
standard film balance was used for the measurements.
Fig. 2 The in-plane excitation (IPE) setup. A set of electrodes, (IDTs), is used to create a
local (electrode distance ≈ 10µm) oscillating electric field in the plane of the lipid monolayer,
which excites a propagating wave by local lateral polarization of the membrane. The substrate
is LiTaO3 being coated with SiO2. To establish a tight contact between the IPE chip and the
lipid monolayer for the detection, the SiO2 is additionally rendered hydrophobic (silanized).
The change in lateral pressure is determined by the force on a Wilhelmy plate dipped into the
monolayer.
Fig. 3 Frequency spectra of the nonlinear pressure change induced by the IPE chip in a DPPG
lipid monolayer (red curve) or pure water for different distances d (black for d=15cm (solid
line) and d=2cm (dotted line)). At lower frequencies, the applied electric field does not affect
the average lateral pressure, being measured by a Wilhelmy plate (see Fig. 1). However, at
certain frequencies (here 10,9MHz and 33MHz), the lateral pressure significantly drops. The
critical frequency (νcrit) on the x-axis can be changed by changing the overall lateral pressure.
We calculate the propagation velocity (c0=λνcrit) from the finger spacing of the stimulating
electrodes and the measured critical frequency νcrit. The inset shows that attenuation is
negligible. In fact, an exponential fit would lead to a decay length of ~1m.
Fig. 4 Frequency spectra of the lateral pressure change, measured for different surface
pressures in a DPPG lipid monolayer at T=14°C. The pressures are chosen to represent the
different phases of the monolayer. The critical frequencies, were the nonlinear pressure
change occurs for the first time, are marked. During maximal compressibility, the critical
frequency is at a clearly lower value, than for the other phase states. Note, that πΔ > 0 above
the maximal compressibility and πΔ < 0 elsewhere. This is correctly predicted by the
assumption of an oscillating density wave (Eq. 8).
Fig. 5 Propagation velocities along the lipid monolayer, measured as described in Fig. 2
[filled squares] and calculated from the adiabatic compressibility as described in the text
[line]. The inserts show the experimentally obtained isothermal compressibility used for the
calculation of c0. (a) shows the velocity curves for DPPG and (b) for a mixture of DPPG and
DPPC (1:10). For high lateral pressures, the agreement between c0 as being predicted from the
12
adiabatic compressibility and the experimentally determined velocity is within less than 5%.
In the phase transition and for the lower pressure regime the deviation increases, but still falls
within a 10% error bar. The origin of the deviation may be due to the fact that the monolayer
becomes less stable at low pressures. Note the larger errors around the maximum in
compressibility. These probably correspond to the increased sensitivity of κs in to oscillations
in density in that region.
Fig. 6 Schematic representation of the important length scales of the system. Both viscous
(δ) and thermal (ξ) penetration depths and the geometry of the system are marked. While the
lateral density wave propagates along the x-y-plane, oscillations both in density and in
temperature decay exponentially along the z-direction. The density length scale (δ) is of the
same order or larger than the thermal penetration depth (ξ) and therefore proposes negligible
dissipation of heat into the bulk water.
13
|
1307.0090 | 1 | 1307 | 2013-06-29T12:16:22 | Self-propelled Janus particles in a ratchet: Numerical simulations | [
"physics.bio-ph",
"cond-mat.soft"
] | Brownian transport of self-propelled overdamped microswimmers (like Janus particles) in a two-dimensional periodically compartmentalized channel is numerically investigated for different compartment geometries, boundary collisional dynamics, and particle rotational diffusion. The resulting time-correlated active Brownian motion is subject to rectification in the presence of spatial asymmetry. We prove that ratcheting of Janus particles can be orders of magnitude stronger than for ordinary thermal potential ratchets and thus experimentally accessible. In particular, autonomous pumping of a large mixture of passive particles can be induced by just adding a small fraction of Janus particles. | physics.bio-ph | physics | Self-propelled Janus particles in a ratchet: Numerical simulations
Pulak Kumar Ghosh1, Vyacheslav R. Misko1,2, Fabio Marchesoni1,3, and Franco Nori1,4
1 CEMS, RIKEN, Saitama, 351-0198, Japan
2Departement Fysica, Universiteit Antwerpen, B-2020 Antwerpen, Belgium
3 Dipartimento di Fisica, Universit`a di Camerino, I-62032 Camerino, Italy and
4 Physics Department, University of Michigan, Ann Arbor, MI 48109-1040, USA
(Dated: June 5, 2018)
Brownian transport of self-propelled overdamped microswimmers (like Janus particles) in a two-
dimensional periodically compartmentalized channel is numerically investigated for different com-
partment geometries, boundary collisional dynamics, and particle rotational diffusion. The resulting
time-correlated active Brownian motion is subject to rectification in the presence of spatial asym-
metry. We prove that ratcheting of Janus particles can be orders of magnitude stronger than for
ordinary thermal potential ratchets and thus experimentally accessible. In particular, autonomous
pumping of a large mixture of passive particles can be induced by just adding a small fraction of
Janus particles.
PACS numbers: 82.70.Dd 87.15.hj 36.40.Wa
Rectification of Brownian motion has been the focus
of a concerted effort, both conceptual [1] and techno-
logical [2], aimed at establishing net particle transport
on a periodic substrate in the absence of external bi-
ases. To this purpose two basic ingredients are required
(Pierre Curie's conjecture): a spatial asymmetry of the
substrate and a time correlation of the (non-equilibrium)
fluctuations, random or deterministic, applied to the dif-
fusing particles. Each particle is assumed to interact with
the substrate via an appropriate periodic potential, also
called ratchet potential. Typically, demonstrations of the
ratchet effect had recourse to external unbiased time-
dependent drives (rocked and pulsated ratchets); recti-
fication induced by time-correlated, or colored, fluctua-
tions (thermal ratchets) seems to be of no practical use,
despite its conceptual interest.
Brownian diffusion in a narrow, corrugated channel
can also be rectified according to Curie's conjecture.
The constituents of a mixture of repelling particles in a
periodically-modulated channel, are pressed against the
channel walls so that their dynamics becomes sensitive
to any asymmetry of the channel compartments (collec-
tive geometric ratchet). Subjected to an a.c. drive ori-
ented along the channel axis, the mixture drifts in the
easy-flow direction, where the average compartment cor-
rugation is the less steep [3], although with much lower
efficiency than in ordinary ratchet potentials. Such a
collective ratchet mechanism has been experimentally ob-
served for a.c. drives and relatively high particle densities
[4, 5], whereas the net current apparently vanishes at low
densities [3]. A simple kinetic equation argument [6, 7]
suggests that rectification of mixtures of repelling parti-
cles, or even single particles, in an asymmetric channel
can also be induced by time-correlated thermal fluctua-
tions, like in thermal ratchets. However, being thermal
ratchets weak in general and (low-density) collective ge-
ometric ratchets less performing than potential ratchets,
demonstration of such an effect seems beyond reach. On
the other hand, rectification of Brownian diffusion by an
internal energy source, like the nonequilibrium fluctua-
tions invoked to power thermal ratchets, is very appeal-
ing: The diffusing particles would harvest kinetic energy
directly from their environment, without requiring any
externally applied field (though unbiased), and transport
would ensue as an autonomous symmetry-directed parti-
cle flow.
To enhance rectification of time correlated-diffusion in
a modulated channel with zero drives, we propose to use
a special type of diffusive tracers, namely of active, or
self-propelled, Brownian particles. Self-propulsion is the
ability of most living organisms to move, in the absence
of external drives, thanks to an"engine" of their own
[8]. Self-propulsion of micro- and nano-particles (arti-
ficial microswimmers) poses a challenge with respect to
their unusual nonequilibrium diffusion properties as well
as their applications to nanotechnology [9]. Recently, a
new type of microswimmers has been synthesized, where
self-propulsion takes advantage of the local gradients that
asymmetric particles can generate in the presence of an
external energy source (self-phoretic effect). Such parti-
cles, called Janus particles [10], consist of two distinct
"faces", only one of which is chemically or physically
active. Such two-faced objects can induce either con-
centration gradients, by catalyzing some chemical reac-
tion on their active surface [11, 12], or thermal gradients,
by inhomogeneous light absorption (self-thermophoresis)
[13, 14] or magnetic excitation (magnetically induced
self-thermophoresis [15]). Moreover, experiments demon-
strated the ability of Janus microswimmers to perform
guided motions through periodic arrays [14] and sepa-
rate colloidal mixtures, due to their selective interaction
with the constituents of the mixture [16].
An active microswimmer gets a continuous push from
the environment, which in the overdamped regime (iner-
tia effects are generally neglected) corresponds to a self-
propulsion velocity v0 with constant modulus v0 and di-
3
1
0
2
n
u
J
9
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
0
9
0
0
.
7
0
3
1
:
v
i
X
r
a
2
.
θ
0τθ/4)t + (v2
rection randomly varying in time with rate τ−1
In a
two-dimensional (2D) boundless suspension, the position
r(t) = (x(t), y(t)) of the microswimmer diffuses accord-
ing to Furth's law
(cid:104)∆r(t)2(cid:105) = 4(D0 + v2
θ /2)(e−2t/τθ − 1), (1)
0τ 2
where ∆r(t) = r(t) − r(0) and D0 is the translational
diffusion constant of a passive particle of the same geom-
etry at a fixed temperature. The mechanisms responsible
for translational and rotational diffusion are not neces-
sarily the same [12, 17] and therefore D0, v0 and τθ can
be treated as independent model parameters. The diffu-
sion law of Eq. (1) is due to the combined action of two
statistically independent 2D Gaussian noise sources [18],
a delta-correlated thermal noise, ξ0(t) and a colored ef-
fective propulsion noise, ξc(t), with correlation functions
(cid:104)ξ0,i(t)(cid:105) = 0, (cid:104)ξc,i(t)(cid:105) = 0, (cid:104)ξ0,i(t)ξ0,j(0)(cid:105) = 2D0δijδ(t),
and (cid:104)ξc,i(t)ξc,j(0)(cid:105) = 2(Dc/τθ)δije−2t/τθ , with i, j = x, y
and Dc = v2
0τθ/4. Correspondingly, the microswimmer
mean self-propulsion path is lθ = v0τθ.
When confined to a constrained geometry,
like a
channel, with compartment size smaller than the self-
propulsion length, lθ, the microswimmer undergoes mul-
tiple collisions with the walls and the confining geometry
comes into play (Knudsen diffusion [19]). Contrary to
standard thermal ratchets in asymmetric potentials [20],
where the strength of the colored noise is kept constant,
here Dc grows linearly with τθ (i.e., the variance of ξc(t)
is set to v2
0). As a consequence, increasing τθ not only
makes geometric rectification effective even in the case of
a single particle, but also enhances the power dissipated
to fuel its self-propulsion. As a result, rectification in
active Brownian ratchets can be so much stronger than
in ordinary thermal ratchets that direct observation be-
comes possible. For instance, our simulations show that
even a small fraction of interacting micro-swimmers suf-
fices to drag along a large mixture of passive particles
(autonomous pumping).
y = v0 sin θ + ξ0,y(t),
Autonomous Janus ratchets. - The rectification of a
Janus particle in a 2D asymmetric channel was simulated
by numerically integrating the Langevin equations [18],
θ = ξθ(t),
x = v0 cos θ + ξ0,x(t),
where ξ0,x(t) and ξ0,y(t) have been defined above and
ξθ(t) is an additional 1D Gaussian noise with (cid:104)ξθ(t)(cid:105) = 0
and (cid:104)ξθ(t)ξθ(0)(cid:105) = 2Dθδ(t), modeling the fluctuations
of the self-propulsion angle θ, measured, say, with re-
spect to the positive channel easy-flow direction. As
(cid:104)cos θ(t) cos θ(0)(cid:105) = (cid:104)sin θ(t) sin θ(0)(cid:105) = (1/2)e−t/Dθ , it
follows immediately that, in the notation of Eq.
(1),
v0 cos θ and v0 sin θ play the role of ξc,x(t) and ξc,y(t), re-
spectively, with τθ = 2/Dθ. The channel compartments
were taken triangular in shape, with length xL, width
yL and pore size ∆ [see inset in Fig. 1(a)]. Through-
out our simulations we kept the compartment aspect ra-
tio r = xL/yL constant with r = 1, and, by rescaling
x → x/κ and y → y/κ, we conveniently set xL = yL = 1.
FIG. 1: (Color online) Rectification of a single pointlike Janus
particle with self-propulsion speed v0 in a triangular channel
with compartment size xL = yL = 1: (a) average velocity
¯v vs.
τθ for channel pore size ∆ = 0.1, different D0 and
sliding (filled symbols) or randomized b.c. (empty symbols).
A dashed line with slope 1 is drawn for reader's convenience.
Inset: (logarithmic) contour plot of the stationary probability
density P (x, y) in a channel compartment; (b) average ¯v vs.
∆ for different D0, sliding b.c. and τθ = 300.
Analogously, we rescaled t → v0t/κ, so as to work with
a constant self-propulsion velocity, v0 = 1. In summary,
the output of our integration code only depends on two
rescaled noise intensities, D0/κv0 and κDθ/v0 (or, equiv-
alently, v0τθ/κ), and one geometric parameter ∆/yL.
The collisional dynamics of a Janus particle at the
boundaries was modeled as follows. The translational
velocity r is elastically reflected, whereas for the coor-
dinate θ we considered two possibilities: (i) frictionless
collisions, θ unchanged. The microswimmer slides along
the walls for an average time of the order of τθ, until
the noise ξθ(t) redirects it toward the center of the com-
partment. The inset in Fig. 1(a) clearly shows the ac-
cumulation of the stationary particle probability density
P (x, y) along the boundaries; (ii) rotation induced by a
tangential friction, θ randomized. This causes the parti-
cle to diffuse away from the boundary, which in general
weakens the rectification effect [see Fig. 1(a)]. Note that
in the case of elastic boundary reflection of both r and
v0, the microswimmer motion would amount to an ordi-
10010210-410-2v0-6-5-4-3-2vp/v0 D0/v0 = = 0.03 (random)D0/v0 = 0 (random) 0.06 0.03 0.02 0(a) pdfMachine Is a pdf writer that produces quality PDF files with ease! Produce quality PDF files in seconds and preserve the integrity of your original documents. Compatible across nearly all Windows platforms, if you can print from a windows application you can use pdfMachine. Get yours now! 0.30.60.90.00.2D0/v0 = v/v0 0.06 0.03 0.02 0.012 0(b) pdfMachine A pdf writer that produces quality PDF files with ease! Produce quality PDF files in seconds and preserve the integrity of your original documents. Compatible across nearly all Windows platforms, simply open the document you want to convert, click "print", select the "Broadgun pdfMachine printer" and that's it! Get yours now! 3
we notice that the curves ¯v(τθ) at low τθ shift upwards
on raising the thermal noise level D0. This effect be-
comes apparent for lθ < xL, that is,
for τθ smaller
than the average time a self-propelled particle takes to
exit a compartment. Under such circumstances, thermal
noise assists the rectification process. Moreover, on tak-
ing the long-time limit of the diffusion law in Eq. (1),
(cid:104)∆r(t)2(cid:105) = 4(D0 + Dc)t, we see that at even smaller self-
propulsion lengths, lθ < 4D0/v0, self-propulsion can be
neglected with respect to thermal fluctuations.
Panel (b) of Fig. 1 illustrates the dependence of ¯v on
the pore size for increasing D0 at large τθ and sliding b.c.
Here the rectification power is suppressed by the thermal
fluctuations and is the highest for large ∆. The first
trend is opposite to that observed for low τθ: now ξ0(t)
helps the Janus particle bypass the compartment corners,
when v0 pushes it in the negative direction [see inset of
panel (a)]. Therefore, the diode-funneling effect exerted
by the triangular compartments can be either enhanced
or weakened by delta-correlated fluctuations, depending
on the regime of self-propulsion. The second trend might
sound counterintuitive, were not that for larger ∆ and
fixed yL the sides of the channel grow shorter and the
particle takes less time to slide along them and through
the exit pore. On the other hand, the backward flow with
negative velocity is blocked mostly at the compartment
corners, regardless of the actual pore size. The moder-
ate ∆ dependence reported in Fig. 1(b) indicates that
our numerical analysis can be safely extended to Janus
swimmers of finite radius.
Finally, in view of practical applications, we tested the
robustness of Janus particle rectification in channels with
variable degrees of asymmetry. In Fig. 2 we varied the
channel geometry by symmetrically shifting the compart-
ment corners by a fixed amount x0 ∈ [0, 0.5) (see inset).
Moreover, we also enlarged the compartments by a scal-
ing factor κ, so as to accommodate for a variety of ex-
perimental set-ups. One immediately sees that the rec-
tification power decreases by only a factor 2 for x0 up
to 0.2 and is insensitive to κ as long as lθ > κxL. For
much larger κ, the particle spends most of its time away
from the (asymmetric) compartment walls and ¯v drops
inversely proportional to κ. Indeed, the rescaled inten-
sity of ξ0(t), D0/κv0, is suppressed with respect to the
rescaled propulsion noise intensity, Dc = κDθ/v0, which
means that for κ → ∞, κ¯v/v0 tends to a constant, that
is, ¯v ∝ κ−1 [see curves for D0 = 0 in Fig. 1(a).]
Janus particles as autonomous pumps.- The remark-
able robustness of the rectification mechanism investi-
gated here lends itself to practical applications. Let
us consider a binary mixture consisting of Nm Janus
microswimmers with large lθ (active particles) and Np
non-self-propelling objects (passive particles). For sim-
plicity, both species are represented by soft interact-
ing disks of radius r0 and repulsive force of modulus
Fi,j = k(2r0 − rij) if rij < 2r0 and Fi,j = 0 other-
FIG. 2: (Color online) Rectification of single Janus particle
with v0 = 1 in asymmetric channels with different geometries.
A typical compartment is sketched in the inset: xL, yL, and
∆ are as in Fig. 1(a), i.e., κ = 1, but the corners are shifted
by x0. The compartment dimensions have then be rescaled
by a magnification factor κ. Main panel: average velocity ¯v
versus κ for τθ = 300, sliding b.c. and different x0. A dashed
line with slope −1 is drawn for reader's convenience.
nary equilibrium Brownian diffusion with finite damping
constant, γ = 2/τθ [21], and rectification would be sup-
pressed.
In Fig. 1 we report our results for the rectification
current, ¯v ≡ (cid:104) x(cid:105) (in units of v0), of a pointlike Janus
particle in a triangular channel with fixed compartment
dimensions and varying τθ, panel (a), and ∆, panel (b).
In panel (a) the pore size was set to ∆ = 0.1 and sev-
eral curves ¯v versus τθ were computed for different D0,
i.e, at different temperatures, and sliding boundary con-
ditions (b.c., filled symbols). At large τθ, microswim-
mer diffusion is of the Knudsen type and rectification is
dominated by self-propulsion; all curves ¯v(τθ) increase
monotonously with τθ until they level off [the weak D0
dependence of such asymptotes is shown in Fig. 1(b)].
Most importantly, we obtained ratios ¯v/v0 in excess of
20%, which means that here the rectification power is or-
ders of magnitude larger than for single-particle thermal
ratchets in an asymmetric potential [2]. Moreover, the
simulation parameter values adopted here, in rescaled
units, are consistent with the corresponding values re-
ported in the experimental literature, see, e.g., Table 1
of Ref. [14]; hence, the possibility of a direct demonstra-
tion of this striking effect.
For the sake of a comparison in panel (a) we also report
two curves ¯v(τθ) obtained by imposing θ randomization
at the boundaries (empty symbols). As expected ¯v/v0 de-
creases as the persistency of the particle self-propulsion
is suppressed by its collisions against the walls. The
thermal fluctuations ξ0(t) further suppress rectification
as they induce more wall collisions and, thus, stronger θ
randomization at the boundaries. In other words, ran-
domized b.c. tend to statistically couple thermal fluctu-
ations and random self-propulsion mechanism. Finally,
10110310-310-10 0.1v/v0 0-1/21/21x0 x0 = 0 0.1 0.2 0.3 pdfMachine A pdf writer that produces quality PDF files with ease! Produce quality PDF files in seconds and preserve the integrity of your original documents. Compatible across nearly all Windows platforms, if you can print from a windows application you can use pdfMachine. Get yours now! wise (Fij and rij denote respectively the pair force and
distance) [18]. Other potentials have also been tested
with qualitatively similar results, namely, (i) active mi-
croswimmers are capable of rectifying their motion even
through a crowd of passive particles, i.e., for packing frac-
tions φ = 2πr2
0Nt/xLyL in excess of 1. [Particles of either
species overlap in average by a length lo = v0/k and can
even pass each other]; (ii) due to their finite size and re-
pulsive interaction with all mixture constituents, even a
low fraction of Janus microswimmers suffices to set the
entire mixture in motion (autonomous pumping).
Autonomous pumping of a binary mixture of Nt =
Nm + Np soft disks was simulated under simplifying as-
sumptions: (i) In the bulk, the microswimmer orienta-
tion randomly changes at time intervals τθ with θ uni-
formly distributed in [0, 2π), which is equivalent to set-
ting Dθ = π2/6τθ; (ii) Collisions with the walls only take
place when the center of the disks hit the boundary; disks
are not elastically repelled by the walls [see inset of Fig.
3(c)]; (iii) Zero thermal fluctuations, D0 = 0, random-
ized b.c. (i.e., θ diffusion is stronger close to the walls
than in the bulk) were assumed for the active microswim-
mer dynamics in order to demonstrate the pumping effect
under the least favorable conditions. Under sliding b.c.
pumping is appreciably stronger (not shown).
Our simulation results are reported in Fig. 3. A few
general properties, largely independent of the model de-
tails, are discussed here. Firstly, the mixture drifts in
the easy-flow (positive) direction, the average velocities
of the active and passive particles being, respectively, ¯vm
and ¯vp. In particular, the rectification power, ¯vp/v0, is
no smaller than reported for potential ratchets [2]. Sec-
ondly, for any given choice of the model parameters, there
exists an optimal Nt where the pumping is the strongest.
This defines an optimal pumping active fraction, typi-
cally with Nm/Nt < 0.1. Thirdly, pumping is clearly
fueled by active microswimmers as proven by ¯vp drop-
ping for lθ (cid:46) xL [panel (a)] and surging with increasing
the interaction constant, k [panel (d)], or the fraction of
active particles (not shown). For harder disks, optimal
pumping occurs at larger τθ because a Janus disk takes
more time (and a longer random path) to cross a com-
partment crowded with more strongly repelling disks. As
for the case of Fig. 1(b), the pore size ∆ is clearly less
important as a control parameter than τθ and k [panel
(d)]. Finally, we notice that for dense mixtures, φ ∼ 1,
the microswimmer velocity ¯vm can reverse sign. This ef-
fect was anticipated for binary mixtures on a ratchet [22]:
passive particles accumulate preferably on the r.h.s. of
the pore, where the cross-section is the largest; passive
particle thus act upon the diluted Janus microswimmers
as an asymmetric repulsive potential barrier, which can
supersede the funneling action of the compartment walls;
hence, the Janus current inversion at high Nt.
To conclude, we stress that the autonomous rectifica-
tion and pumping effects discussed here apply to biologi-
4
FIG. 3: (Color online) Rectification of a binary mixture made
of Nm = 4 Janus particles with self-propulsion speed v0,
D0 = 0, and Np = Nt − Nm passive particles. All parti-
cles are modeled as elastically repelling soft disks of radius
r0 = 0.05; channel compartments are triangular with dimen-
sions xL = yL = 1. In the inset, active and passive particles
are represented by (blue) filled and (red) empty circles, re-
spectively. Panels (a) and (b): average rectification velocity
(a) of passive, ¯vp and (b) active particles, ¯vm, versus Nt for
∆ = 0.1 and different interaction constant, k [see legend in
(b)]; (c) ¯vp vs. Nt for ∆ = 0.1 and different τθ; (d) ¯vp vs. τθ
for Nt = 72 and different ∆ and k. Note that lo = v0/k is a
measure of the disk overlap, so that τθk = l/lo.
cal and artificial swimmers regardless of their propulsion
mechanism, and not only to especially fabricated Janus
particles. For instance, the autonomous robots of Ref.
[23] are laser driven with propulsion speed v0 and about
10 µm across, which means that their translational dif-
fusion is negligible. The diffusion of their propulsion ori-
entation is due to the scattering by spatial disorder, that
is, τθ ∼ l/v0, where l is a disorder correlation length. In
cellular systems, propulsion of macro-biomolecules can
be fueled by the "power-stroke" associated with the hy-
drolysis of ATP in suspension. In that case, v0 ∼ δr/τθ,
where δr is the net displacement produced by a single
power-stroke and τ−1
coincides with the ATP hydrolysa-
tion rate in the vicinity of the biomolecule [19].
θ
20406080022040608003k/v0 =vp/v0 (x 103)vm/v0 (x 100)Nt 10 20 40 60(a)(b) Nt pdfMachine - is a pdf writer that produces quality PDF files with ease! Get yours now! "Thank you very much! I can use Acrobat Distiller or the Acrobat PDFWriter but I consider your product a lot easier to use and much preferable to Adobe's" A.Sarras - USA 255075010202 1.5 1 0.5 10 5 2.5Ntv0 tq =D, tqk = 0.1 ,10 0.05,10 0.01,10 0.05, 20 0.05, 40 v0tq'vp/v0 (x 103)'vp/v0 (x 103)(c)(d)Acknowledgements
We thank RICC for computational resources. PKG
acknowledges financial support from JSPS through fel-
lowship No. P11502. VRM acknowledges support from
the Odysseus Program of the Flemish Government and
FWO-VI. FM acknowledges partial support from the Eu-
ropean Commission, grant No. 256959 (NanoPower). FN
was supported in part by the ARO, JSPS-RFBR contract
No. 12-02-92100, Grant-in-Aid for Scientific Research
(S), MEXT Kakenhi on Quantum Cybernetics, and the
JSPS via its FIRST program.
[1] P. Reimann, Phys. Rep 361, 57 (2002).
[2] P. Hanggi and F. Marchesoni, Rev. Mod. Phys. 81, 387
(2009); P. Hanggi, F. Marchesoni, F. Nori, Brownian mo-
tors, Annalen der Physik 14, 51 (2005).
[3] J.F. Wambaugh, C. Reichhardt, C.J. Olson, F. March-
esoni, and F. Nori, Phys. Rev. Lett. 83, 5106 (1999).
[4] J.E. Villegas, S. Savel'ev, F. Nori, E.M. Gonzalez, J.V.
Anguita, R. Garc´ıa, and J.L. Vicent, Science 302, 1188
(2003); Y. Togawa, K. Harada, T. Akashi, H. Kasai,
T. Matsuda, F. Nori, A. Maeda, A. Tonomura, Phys.
Rev. Lett. 95, 087002 (2005); C.C. de Souza Silva, J.
Van de Vondel, M. Morelle, and V.V. Moshchalkov, Na-
ture 440, 651 (2006); N.S. Lin, T.W. Heitmann, K. Yu,
B.L.T. Plourde, and V.R. Misko, Phys. Rev. B 84,
144511 (2011).
[5] D. Cole, S. Bending, S. Savel'ev, A. Grigorenko, T.
Tamegai, and F. Nori, Nature Materials 5, 305 (2006).
[6] R. Zwanzig, J. Phys. Chem. 96, 3926 (1992).
[7] for a review see P.S. Burada, P. Hanggi, F. March-
esoni, G. Schmid, and P. Talkner, ChemPhysChem 10,
45 (2009).
[8] E.M. Purcell, Am. J. Phys. 45, 3 (1977).
5
[9] F. Schweitzer, Brownian Agents and Active Particles
(Springer, Berlin, 2003); P. Romanczuk, M. Bar, W.
Ebeling, B. Lindner, and L. Schimansky-Geier, Eur.
Phys. J. Special Topics 202, 1 (2012).
[10] S. Jiang and S. Granick (Eds.), Janus Particle Synthesis,
Self-Assembly and Applications (RSC Publishing, Cam-
bridge, 2012); A Walther and A.H.E. Muller, Chem. Rev.
(2013) doi: 10.1021/cr300089t
[11] W.F. Paxton, S. Sundararajan, T.E. Mallouk, A. Sen,
Angew. Chem. Int. Ed. 45, 5420 (2006).
[12] J.G. Gibbs and Y.-P. Zhao, Appl. Phys. Lett. 94, 163104
(2009); J.R. Howse, R.A.L. Jones, A.J. Ryan, T. Gough,
R. Vafabakhsh, R. Golestanian, Phys. Rev. Lett. 99,
048102 (2007).
[13] H.R. Jiang, N. Yoshinaga, and M. Sano, Phys. Rev. Lett.
105, 268302 (2010).
[14] G. Volpe, I. Buttinoni, D. Vogt, H.-J. Kummerer, and C.
Bechinger, Soft Matter 7, 8810 (2011).
[15] L. Baraban, R. Streubel, D. Makarov, L. Han, D. Kar-
naushenko, O.G. Schmidt, and G. Cuniberti, ACS Nano
7, 1360 (2013).
[16] W. Yang, V.R. Misko, K. Nelissen, M. Kong, and F.M.
Peeters, Soft Matter 8, 5175 (2012).
[17] S. van Teeffelen and H Lowen, Phys. Rev. E 78,
020101(R) (2008).
[18] Y. Fily and M.C. Marchetti, Phys. Rev. Lett. 108,
235702 (2012).
[19] H. Brenner and D. A. Edwards, Macrotransport Processes
(Butterworth-Heinemann, New York, 1993)
[20] R. Bartussek, P. Reimann, and P. Hanggi, Phys. Rev.
Lett. 76, 1166 (1996).
[21] P. K. Ghosh, P. Hanggi, F. Marchesoni, F. Nori, and G.
Schmid, EPL 98, 50002 (2012); Phys. Rev. E 86, 021112
(2012).
[22] S. Savel'ev, F. Marchesoni, and F. Nori, Phys. Rev. Lett.
91, 010601 (2003); 92, 160602 (2004).
[23] A. B´uz´as, L. Kelemen, A. Mathesz, L. Oroszi, G. Vizs-
nyiczai, T. Vicsek, and P. Ormos, Appl. Phys. Lett. 101,
041111 (2012).
|
1205.2867 | 1 | 1205 | 2012-05-13T14:21:50 | Theory of single-molecule experiments in the overstretching force regime | [
"physics.bio-ph",
"cond-mat.soft"
] | We present a statistical mechanics analysis of the finite-size elasticity of biopolymers, consisting of domains which can exhibit transitions between more than one stable state at large applied force. The constant-force (Gibbs) and constant-displacement (Helmholtz) formulations of single molecule stretching experiments are shown to converge in the thermodynamic limit. Monte Carlo simulations of continuous three dimensional polymers of variable length are carried out, based on this formulation. We demonstrate that the experimental force-extension curves for short and long chain polymers are described by a unique universal model, despite the differences in chemistry and rate-dependence of transition forces. | physics.bio-ph | physics |
03/2012-PRL
Theory of single-molecule experiments in the overstretching force regime
Fabio Manca1, Stefano Giordano2, Pier Luca Palla2,3, Fabrizio Cleri2,3, and Luciano Colombo1
1Dipartimento di Fisica, Universit`a di Cagliari, Cittadella Universitaria, 09042 Monserrato (Ca), Italy
2Institut d'Electronique, Micro´electronique et Nanotechnologie (CNRS UMR 8520) and
3Universit´e de Lille I, 59652 Villeneuve d'Ascq, France
(Dated: November 16, 2018)
We present a statistical mechanics analysis of the finite-size elasticity of biopolymers, consist-
ing of domains which can exhibit transitions between more than one stable state at large applied
force. The constant-force (Gibbs) and constant-displacement (Helmholtz) formulations of single
molecule stretching experiments are shown to converge in the thermodynamic limit. Monte Carlo
simulations of continuous three dimensional polymers of variable length are carried out, based on
this formulation. We demonstrate that the experimental force-extension curves for short and long
chain polymers are described by a unique universal model, despite the differences in chemistry and
rate-dependence of transition forces.
PACS numbers: 87.15.By, 83.10.Nn, 87.15.He
Dynamic force spectroscopy by means of the atomic-
force microscope (AFM), laser- or magnetic-tweezers ap-
paratus, or the biomembrane force probe, allows the di-
rect probing of the elasticity of individual molecules, and
as such has rapidly become a mainstay of biophysical re-
search [1 -- 4]. These mechanical devices are quite different
from one another, one prominent difference being their
equivalent stiffness, in the range of 10−4 − 1 pN/nm for
tweezers, vs. 10 − 105 pN/nm for the AFM [4]. The typ-
ical experiment is a mechanically-induced unfolding of a
biological polymer made of N domains, e.g. a polysac-
charide such as dextran [5], a protein such as titin [6],
a DNA or RNA strand [7], and so on. As a function
of increasing force levels different mechanical response
regimes are observed, beginning with the entropic unfold-
ing of the polymer chain (now well understood in terms
of simple worm-like chain (WLC) or freely-jointed chain
(FJC) models [8]); to the linear-elastic extension of the
straightened chain; to the so-called overstretching, typ-
ically interpreted as a conformational transformation of
the domain geometry; up to the eventual fracturing of
the polymer.
In this Letter we provide a robust statistical mechan-
ics foundation to the interpretation of the overstretching
regime, which we describe in terms of the internal dy-
namics of a chain of two -- state systems undergoing a con-
formational transformation, as described by the double-
well potential in Fig.1. For the sake of argument we
call 'folded' and 'unfolded' the two conformations; how-
ever the transformation occurs, more generally, between
two principal local minima of the domain free-energy hy-
persurface (e.g., for DNA it could as well represent the
melting transition [11]). Then, we firstly develop a theo-
retical model describing experiments at constant applied
force (a realization of Gibbs ensemble statistics) and we
show that the conformational change must occur simul-
taneously for all the domains at a given threshold force.
experiments performed at
On the other hand,
constant -- displacement are a realization of the Helmholtz
ensemble of statistical mechanics. In our previous work
[9], we showed that the outcome of the two types of ex-
periment converge in the thermodynamic limit of infinite
chain length, N → ∞. In practice, real experiments al-
ways fall inbetween these two ideal extremes. Therefore,
here we focus on the intermediate cases described by fi-
nite values of the kc/k ratio, k and kc being the equivalent
spring constant (i.e., stiffness) of the domain and of the
pulling device, respectively. We demonstrate by means
of Monte Carlo simulations that the typical 'sawtooth'
pattern [6], observed for the unfolding of large protein
domains (such as the Ig units in titin), and the 'plateau'
or kink [5, 7], observed in the overstretching of DNA and
polysaccharides (e.g. dextran), have a common origin
in the size-dependence of the polymer response to the
external force, the plateau shape being attained in the
limit of large N . On the same grounds, at a fixed num-
xf
V (x)
M
∆E
0
xu
xf
xf
xf
x1
x0
x2
xu
x
FIG. 1: Potential energy function with an energy bar-
rier. Folded and unfolded configurations of the domains are
schematically represented.
ber N of domains, the transition from the 'plateau' to
the 'sawtooth' response is recovered for increasing values
of kc/k. Notably, such a behavior of the force-extension
curves is universal with respect to the specification of
any additional parameters, such as chemical, structural
or mechanical constants of the domains.
We work out a simple model containing the minimal
ingredients fully describing the overall complex behavior
of a polymer chain.
It consists of an N -domain, non-
branched chain clamped at one end, able to describe con-
formational transitions across an energy barrier. The in-
ternal state of each domain is described by a potential
energy V (x) which exhibits two minima corresponding
to the lengths x = xf (folded conformation) and x = xu
(unfolded conformation), connected via an energy barrier
M at x = x0 (see Fig.1). The energy is written as a C2
piecewise function, constructed by imposing continuity
and differentiability at the joining points x1 and x2:
V (x) =
0 < x < x1
2 k(x − x0)2 + M x1 < x < x2
1
2 k(x − xf )2
− 1
1
2 k(x − xu)2 + ∆E x > x2
(1)
For chosen values of the lengths xf and xu, the domain
spring constant k, and the energy difference ∆E between
the two conformations, the other parameters are simply
given by: δ = xu − xf , x0 = (xu + xf )/2 + 2∆E/(kδ),
M = (k/4)[δ/2 + 2∆E/(kδ)]2, x1 = xf + δ/4 + ∆E/(kδ)
and x2 = xu − δ/4 + ∆E/(kδ). Therefore, this model
properly gives a barrier with xf < x0 < xu only for
∆E ≤ kδ2/4.
Upon application of a constant force f to the end
of the polymer identified by the position vector ~rN =
(xN , yN , zN ) (the other end being fixed in the origin),
the statistics of the fluctuating chain is a realization of
the Gibbs ensemble [9]. The ensemble partition func-
eh/kB T dqN dpN , with
tion is given by Zf (f, T ) =R RΓN
ΓN = ℜ6N . The augmented Hamiltonian eh includes the
classical kinetic energy of the domains with mass m, their
total potential energy, and a term, −f zN , describing the
applied force along the z-axis [9]. In the framework of
the present minimal model, the partition function can be
explicitly calculated as:
e−
Zf (f, T ) =(cid:18) 2πm
β (cid:19)3N/2(cid:18) 2π
βf(cid:19)N
(2)
× {Π (βk, βf, xf , 0, x1) + e−βM Π (−βk, βf, x0, x1, x2)
with β=(kBT )−1 and
+e−β∆EΠ (βk, βf, xu, x2, +∞)(cid:9)N
Π (α, γ, x0, a, b) = 2Z b
2 (x−x0)2
xe−
α
a
sinh (γx)dx
(3)
The extension r at a given force is obtained from the
partition function as r = kBT (∂ log Zf /∂f ). Since the
2
50
40
30
20
10
0
0
k
∆ E
1
3
normalized extension
2
4
e
c
r
o
f
d
e
z
i
l
a
m
r
o
n
FIG. 2: (color online) Force-extension curves for the Gibbs
ensemble: normalized force f /fβ vs normalized extension
r/(N xf ). The black solid lines correspond to different values
of the energy ∆E=0, 10, 20, 30, 40, 50 kBT (increasing val-
ues from the bottom up) for a fixed spring constant k = 2000
kBT /(nm)2. The blue dashed lines correspond to different
values of the spring constant k=10, 15, 30, 100 kBT /(nm)2
(increasing values from the right to the left) for a fixed value
of the energy barrier ∆E = 30 kBT .
extension is linearly dependent on N , the data for chains
of different lengths can be scaled to a single curve upon
diving by N .
Figure 2 shows the results of the normalized force-
extension curves, f /fβ (where f −1
β = βxf ) in terms of
r/(N xf ) for different values of the energy barrier ∆E=0,
10, 20, 30, 40, 50 kBT (black solid lines) at a fixed value
of k=2000 kBT /(nm)2, and for different values of the
spring constant k (blue dashed lines) at a fixed value of
∆E=30 kBT . Both sets of curves display a force plateau
at f ≃ ∆E/δ, for any ∆E > 0, with a normalized width
equal to δ. In our model, the plateau indicates a tran-
sition in the polymer conformation, meaning that for
f < ∆E/δ each domain is found in the folded confor-
mation at x=xf , while for f > ∆E/δ domains are in
the unfolded conformation at x=xu.
i.e., the ensemble
of domains respond cooperatively to the external force.
Notably, the value of the plateau force inducing the con-
formation transition does not depend on the spring con-
stant, k, nor on the temperature. Such a result is readily
interpreted in the framework of the Bell -- Kramers theory
[10], as the threshold value of force necessary to make the
unfolding rate equal to the (reverse) folding one, i.e. low-
ering the difference ∆E to zero. For example, in the case
of dsDNA, which displays a plateau at f =65 pN with a
δ ≈ 2A, our criterion gives a value ∆E=3.5 kBT , in fair
agreement with available experimental data [11, 12]. A
similar plateau was observed for other long chain poly-
mers, such as dextran with N =275, xf =0.5 nm, xu=0.56
nm, ∆E = 13.2 kBT [5], for which the simple criterion
f ≈ ∆E/δ gives plateau forces in the range of ≈900 pN,
as indeed observed [13].
While Gibbs ensemble statistics are sampled with a
constant applied force, a dual situation can be realized
e
c
r
o
f
d
e
z
i
l
a
m
r
o
n
50
40
30
20
10
0
0
50
40
30
20
10
e
c
r
o
f
d
e
z
i
l
a
m
r
o
n
0
0
N = 4
kc
1
2
normalized extension
N = 4, 30, 300
1
2
normalized extension
3
3
FIG. 3: (color online) Monte Carlo force-extension curves at
T =293 K, for: (top panel) different decreasing values of the
device spring constant kc =5, 2, 1, 0.5, 0.01 kBT /(nm)2 (from
the top down) and N =4; (lower panel) increasing number of
domains N = 4, 30, 300 with kc = 2 kBT /(nm)2 (bottom
panel). The red dashed line corresponds to the Gibbs ensem-
ble. The remaining parameters are ∆E = 30 kBT , xf = 2.5
nm, xu = 3xf and k = 100 kBT /nm2.
by imposing the extension, i.e. by controlling the poly-
mer end-to-end distance. The statistics of the fluctu-
ating polymer in this latter scheme is a realization of
the Helmholtz ensemble. As shown in Ref.[9], the cor-
responding partition function Zr cannot be written in
closed form and, as opposed to the Gibbs case, the cor-
responding extension r is non-linearly dependent on N .
However, we showed that the partition functions in the
two ensembles are formally related via a Laplace trans-
form, and we demonstrated [9] that they lead to a com-
mon force-extension curve in the thermodynamic limit.
It should be noted that any AFM or tweezers exper-
iment falls in an intermediate regime between the two
ideal extremes, of purely constant -- force or constant --
extension, since either constraint on the terminal domain
of the chain is mediated by a mechanical device (such
as the AFM cantilever, or the laser -- bound microsphere,
plus a molecular spacer providing adhesion). The device
is characterized by its own effective elastic constant kc,
which is coupled in series to the chain of domain springs
k. In the limit of a soft device, kc/k →0, the statistics of
the coupled system reduces to the Gibbs ensemble for the
isolated molecule fluctuating under a constant force. On
3
the other hand, for a very stiff device, kc/k → ∞, one re-
covers the Helmholtz ensemble for the isolated molecule
held at a fixed extension by the fluctuating force [14].
To describe such a situation, we adopt a Monte Carlo
(MC) numerical approach, simulating the stretching of
the chain produced by a device with a proper adjustable
elastic stiffness.
In Figure 3, top panel, we report the results of the MC
simulations at T =293 K, for decreasing values of the kc/k
ratio, from 0.05, that is well within the Helmholtz statis-
tics regime, down to 1 × 10−4, i.e., approaching Gibbs
ensemble statistics. The remaining parameters are set
to N =4, ∆E=30 kBT , xf =2.5 nm, xu=3xf and k=100
kBT /nm2, which can be considered representative of a
medium -- sized, multi -- domain chain protein. At large val-
ues of kc/k, the domains exhibit a sequence of indepen-
dent conformational transitions to the unfolded configu-
ration, generating a series of N peaks (sawtooth pattern)
which closely resemble the experimental results obtained
for short chains (e.g., a titin fragment with N =8, xf =4
nm, xu=32 nm, ∆E=11.1 kBT [13]). For kc/k → 0
the peak-to-valley width, ∆f , of the sawtooth shrinks
and the curve approaches the kc=0 cooperative plateau
of Gibbs statistics.
In substantial agreement with this
finding, pulling experiments on native titin by means of
optical tweezers [15], having a very small equivalent kc
compared to the AFM one, do not reveal the sawtooth
pattern, but rather a smooth, monotonic branch reminis-
cent of the horizontal plateau.
On the other hand, a similar asymptotic trend is ob-
served (Fig. 3, bottom panel) when the chain length, i.e.
the number of domains, is increased, at a fixed value of
kc/k. As N increases, the width ∆f is decreased un-
til, at a large enough N , the force-extension curves ap-
proach again the plateau curve of the Gibbs ensemble.
It is worth noting that a similar trend was observed in
experiments performed on native titin, comprising sev-
eral hundreds of Ig domains, for which the width ∆f was
of the order of 80 pN [6], compared to the much shorter
8-monomer titin, for which ∆f > 200 pN. The experi-
ments performed on dextran, a long polysaccharide with
N = 275, [5, 13] whose response to the applied force
shows a plateau closer to the typical DNA-like behav-
ior, can also be rationalized on this basis. In summary,
we proved that the macroscopically different behavior
of small-N polymers (such as titin) vs.
long polymers
(such as dextran, DNA), as well as experiments done on
a same polymer but with devices having widely different
stiffness, can be interpreted with the very same unifying
model, interpolating between the two extremes of pure
Gibbs or Helmholtz statistics.
As observed by several authors, each branch of the
sawtooth pattern can be nicely fitted by a sequence of
FJC, or WLC curves (see Fig.4, left panel, dashed lines)
with a proper value of the persistence length, up to the
unfolding of each domain (see e.g. titin [6], spectrin [16],
40
30
e
c
r
o
f
d
e
z
i
l
a
m
r
o
n
20
10
0
∆ f / fβ
θ
0
1
normalized extension
2
)
β
f
/
f
∆
(
g
o
l
2
1.5
1
0.5
0
−0.5
k
)
θ
(
n
a
t
3
1
2
log(N)
3
4
0
−50
−100
−150
−200
−250
0
4
k
5
kc
10
FIG. 4: (color online) Left panel: definition of ∆f and θ for a typical force-extension curve with N = 4. Dashed lines for
the growing branches fitted to FJC with increasing contour length. Central panel: plot of log(∆f /fβ) vs.
log(N ) for kc=1
kBT /(nm)2 and k/kc=20, 30, 100 (blue, red and black line, respectively). Right panel: plot of tan(θ) vs. kc for N =4, and the
same k values. Remaining parameters ∆E = 30 kBT , xf = 2.5 nm, xu = 3xf and T =293 K.
fibronectin [17], synaptotagmin [18]). Beyond this point,
the force relaxes to a smaller value, until the next curve is
met and the force can start rising again upon increasing
displacement.
Since the physical origin of the growing branch of the
curves is well understood on the basis of FJC or WLC
models, we analyzed the decreasing branch, as identi-
fied by the common width ∆f and angle θ in Fig.4 (left
panel) which were extracted from our MC simulations as
a function of N and kc/k.
By looking at Fig.
4 (center), the peak-to-valley
width shows a power-law decrease with the chain length,
∆f ∼ N −α, the exponent α = 1.3 being remarkably
independent on the kc/k ratio. This finding indicates
that attainment of the thermodynamic limit is mainly
dictated by the thermal force scale, fβ, and to a much
lesser extent by other structural and chemical details of
the polymer.
It is worth noting that the value of the
exponent is in agreement with previous results on mono-
stable FJC and WLC models with extensible bonds [9].
The last plot on the right of Fig. 4 reports the behav-
ior of tan(θ) as a function of the device stiffness, kc. The
observed linear dependence is another remarkable result,
completely describing the transition between the two ex-
tremes (Gibbs and Helmholtz ensembles), while taking
into account all the intermediate cases. For kc/k → ∞
we have tan(θ) → −∞ or, equivalently, θ → π/2.
In other words, the decreasing branches of the force-
extension curve must be exactly vertical in the case of
the Helmholtz ensemble. Notably, both the values of ∆f
and θ are fully prescribed, i.e. the entire shape of the
force-extension curve is uniquely defined, once the free
parameters of the model are specified.
In conclusion, we described the statistical mechanics
of chain polymers composed by domains with two stable
states, subject to a pulling force by a molecular-scale me-
chanical device. We showed that for short chain length,
or large stiffness of the device, the domain response is
uncorrelated and originates the typical sawtooth force-
extension curve observed in many experiments. On the
other hand, upon increasing chain length, or vanishing
device stiffness, the response is cooperative and results
in the plateau-like curve, also observed in other experi-
ments. Despite the simplicity of the model, such a frame-
work provides a unified picture for such apparently con-
trasting experimental situations.
F.M. acknowledges the University of Cagliari for the
extended visiting grant and the IEMN, University of Lille
I, for the kind hospitality. L.C. acknowledges financial
support by RAS under the project 'Modellizzazione Mul-
tiscala della Meccanica dei Materiali Complessi (M4C)'.
[1] F. Ritort, J. Phys.: Condens. Matter 18, R531 (2006)
[2] F. Cleri, Sci. Model. Simul. 15, 369 (2008)
[3] K. R. Chaurasiya et al. Phys. Life Rev. 7, 299 (2010)
[4] K. C. Neuman, A. Nagy, Nature Meth. 5, 491 (2008)
[5] M. Rief et al., Science 275, 28 (1997).
[6] M. Rief et al., Science 276 1109 (1997).
[7] S. M. Smith, Y. Cui, C. Bustamante, Science 271 795
(1996)
[8] J. F. Marko, E. D. Siggia, Macromolecules 28, 8759
(1995).
[9] F. Manca et al., J. Chem. Phys. 136, 154906 (2012)
[10] G. I. Bell, Science 200, 618 (1978)
[11] P. Cluzel et al., Science 271, 792 (1997)
[12] A. Ahsan, J. Rudnick, R. Bruinsma, Biophys. J. 74, 132
(1998)
[13] M. Rief, J. M. Fernandez, H. E. Gaub, Phys. Rev. Lett.
81 4764 (1998)
[14] H. J. Kreuzer, S. H. Payne, Phys. Rev. E 63, 021906
(2001)
[15] M. S. Kellermayer et al., Science 276, 1112 (1997)
[16] M. Rief et al., J. Mol. Biol. 289 553 (1999)
[17] A. F. Oberhauser et al., J. Mol. Biol. 319 433 (2002)
[18] M. Carrion-Vazquez et al., Prog. Biophys. Mol. Biol. 74
63 (2000)
|
1601.06730 | 2 | 1601 | 2016-06-16T17:16:32 | Derivation of Hodgkin-Huxley equations for a Na+ channel from a master equation for coupled activation and inactivation | [
"physics.bio-ph"
] | The Na+ current in nerve and muscle membranes may be described in terms of the activation variable m(t) and the inactivation variable h(t), which are dependent on the transitions of S4 sensors of each of the Na+ channel domains DI to DIV. The time-dependence of the Na+ current and the rate equations satisfied by m(t) and h(t) may be derived from the solution to a master equation which describes the coupling between two or three activation sensors regulating the Na+ channel conductance and a two stage inactivation process. If the inactivation rate from the closed or open states increases as the S4 sensors activate, a more general form for the Hodgkin-Huxley expression for the open state probability may be derived where m(t) is dependent on both activation and inactivation processes. The voltage dependence of the rate functions for inactivation and recovery from inactivation are consistent with the empirically determined expressions, and exhibit saturation for both depolarized and hyperpolarized clamp potentials. | physics.bio-ph | physics | Derivation of Hodgkin-Huxley equations for a Na+ channel from
a master equation for coupled activation and inactivation
S. R. Vaccaro
Department of Physics, University of Adelaide, Adelaide, South Australia,
5005, Australia
[email protected]
Abstract
The Na+ current in nerve and muscle membranes may be
described in terms of the activation variable m(t) and the in-
activation variable h(t), which are dependent on the transitions
of S4 sensors of each of the Na+ channel domains DI to DIV.
The time-dependence of the Na+ current and the rate equations
satisfied by m(t) and h(t) may be derived from the solution to
a master equation which describes the coupling between two or
three activation sensors regulating the Na+ channel conductance
and a two stage inactivation process.
If the inactivation rate
from the closed or open states increases as the S4 sensors acti-
vate, a more general form for the Hodgkin-Huxley expression for
the open state probability may be derived where m(t) is depen-
dent on both activation and inactivation processes. The voltage
dependence of the rate functions for inactivation and recovery
from inactivation are consistent with the empirically determined
expressions, and exhibit saturation for both depolarized and hy-
perpolarized clamp potentials.
6
1
0
2
n
u
J
6
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
0
3
7
6
0
.
1
0
6
1
:
v
i
X
r
a
1
INTRODUCTION
The opening and subsequent inactivation of Na+ channels and the ac-
tivation of K+ channels generate the action potential in nerve and muscle
membranes [1]. The time-dependence of the Na+ current in the squid axon
may be described in terms of the expression m(t)3h(t) where the activation
variable m(t) and inactivation variable h(t) satisfy the rate equations
dm
dt
= αm − (αm + βm)m,
dh
dt
= αh − (αh + βh)h,
(1)
(2)
and αm, βm, αh, and βh are voltage dependent rate functions for activation
and inactivation transitions across the membrane.
The Hodgkin-Huxley (HH) description of the Na+ current is equivalent
to an 8-state master equation where three independent voltage sensors may
activate and open the channel and independent inactivation may occur from
each of the closed or open states [2, 3, 4]. Although this master equation is
not consistent with the measurement of an almost zero Na+ current during
repolarization of an inactivated channel, by assuming that the deinactivation
rate to the open state is zero but the rate to closed states increases as the
S4 sensors deactivate [5], and that the rate functions satisfy microscopic
reversibility, the model provides a good description of the recovery from
inactivation, and the Na+ current during a depolarizing clamp [6].
In this paper, it is shown that the Hodgkin-Huxley expression for the
Na+ current and the rate equations for activation and inactivation may be
derived from a master equation, which describes the coupling between two
or three activation sensors regulating the Na+ channel conductance and
a two-stage inactivation process. For a Na+ channel with two activation
sensors, where the deactivation rate during repolarization is slower between
inactivated states than between closed or open states, only four of the terms
of the solution to a six state master equation contribute to the dynamics,
and if the activation sensors are independent, the open state O(t) may be
expressed as m(t)2h(t). The voltage dependence of the rate functions for
inactivation and recovery from inactivation have a similar form to empiri-
cal expressions for Na+ channels [1, 5], and in particular, the exponential
variation exhibits saturation for both depolarized and hyperpolarized clamp
potentials.
2
VOLTAGE CLAMP OF A Na+ CHANNEL WITH TWO AC-
TIVATION SENSORS
The Na+ channel protein is comprised of four domains DI to DIV, each
containing an S4 segment with positively charged residues at every third
position [3]. Based on voltage clamp fluorometry, it has been shown that, in
response to membrane depolarization, the transverse motion of the charged
S4 segments of the Na+ channel domains DI to DIII is associated with ac-
tivation, whereas the slower movement of DIV S4 is correlated with the
binding of an intracellular hydrophobic motif that blocks the flow of ions
through the inner mouth of the pore [7]. This may occur for small depolar-
izations when the ion channel is usually closed (closed-state inactivation) or
for larger depolarizations when the S4 segments of the domains D1 to D3
are activated (open-state inactivation).
However, during repolarization of an inactivated Na+ channel, the OFF
gating charge has a fast component which may be attributed to the motion
of the DI and DII S4 segments, and a slow component, the "immobilized"
portion, that is generated by the conformational changes of the DIII and
DIV S4 segments [8, 9]. For an inactivation modified mutant of the human
heart Na+ channel, it has been estimated that the DIV S4 sensor contributes
approximately 30 % to the OFF charge, and approximately 20 % may be
attributed to the DIII S4 sensor which is only immobilized when the inac-
tivation gate is intact. The slow component of the OFF gating charge has
the same time-course as the Na+ channel recovery from inactivation, and
therefore, the rate-limiting step is the motion of the DIV S4 sensor and not
the unbinding of the inactivation gate [10].
In order to account for the effect of double-cysteine mutants of S4 gating
charges on the ionic current of the bacterial Na+ channel NaChBac, it has
been proposed that at least two transitions are required during the activation
of each voltage sensor [11]. This conclusion is consistent with an earlier
result that cross-linking a DIV S4 segment from the extracellular surface
inhibits inactivation during membrane depolarization whereas cross-linking
the same segment from the inside inhibits activation of the Na+ channel, and
therefore, the DIV S4 sensor translocates across the membrane in two stages
[12, 13]. The measurement of currents for charge neutralized segments in
each domain of the Na+ channel gives additional support to the conclusion
that the two-stage activation of the DIV S4 sensor is correlated with ion
channel inactivation [6].
In this section, we assume that the activation of two voltage sensors
regulating the Na+ channel conductance (DIII S4 and the S4 segment of
either the DI or DII domains) is coupled to a two-stage inactivation process
3
(see Fig. 1), and therefore, the kinetics may be described by a master
equation where the occupation probabilities of the closed states C1, C2, A1
and A2, the open states O and A3, and the inactivated (or blocked) states
B1, B2 and B3 are determined by
dC1
dt
dC2
dt
dO
dt
dA1
dt
dA2
dt
dA3
dt
dB1
dt
dB2
dt
dB3
dt
= −(αi1 + αC)C1(t) + βC C2(t) + βi1A1(t)
= −(αi2 + αO + βC)C2(t) + αCC1(t) +
βOO(t) + βi2A2(t)
= αOC2(t) − (βO + αi3)O(t) + βi3A3(t)
= αi1C1(t) − (αA1 + βi1 + γi1)A1(t)
+δi1B1(t) + βA1A2(t)
= αi2C2(t) − (αA2 + βA1 + βi2 + γi2)A2(t)
+δi2B2(t) + αA1A1(t) + βA2A3(t)
= αi3O(t) − (βA2 + βi3 + γi3)A3(t)
+δi3B3(t) + αA2A2(t)
= γi1A1(t) − (αB1 + δi1)B1(t) + βB1B2(t)
= γi2A2(t) + αB1B1(t) + βB2B3(t) −
(αB2 + βB1 + δi2)B2(t),
= γi3A3(t) + αB2B2(t) − (βB2 + δi3)B3(t).
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
The master equation may be derived from a Smoluchowski equation
applied to the resting and barrier regions of an energy landscape for each
of the S4 sensors in the domains DI to DIV [14, 15]. The translocation of
the S4 segment through the gating pore for Na+ (or K+) channels requires
sufficient energy to overcome several barriers that are dependent on the
Coulomb force between positively charged residues on the S4 sensor and
negatively charged residues on neighboring helices, the dielectric boundary
force, the electric field between internal and external aqueous crevices, and
hydrophobic forces [16]. It is assumed that the transition rates for each stage
of inactivation are dependent on single barrier activation, and therefore, are
proportional to exp(-U) where U is the voltage dependent height of the
barrier [17]. However, if the Na+ channel S4 sensors of the DI, DII or DIII
4
domains are activated in two stages, the rate functions αm and βm may be
approximated by two-state expressions [18].
In order to simplify the solution of Eqs. (3) to (11), it is initially assumed
that
αik = αi1, βik = βi1, γik = γi1,
(12)
for each k, and to ensure that the Na+ current recovers from inactivation
when the S4 sensors that regulate Na+ conductance deactivate, it is further
assumed that
δi1 > δi2 > δi3 ≈ 0.
(13)
From microscopic reversibility or the principle of detailed balance, the prod-
uct of the transition rates in the clockwise and anticlockwise directions are
equal [3], and we may write
δ21
δ32
αB1
βB1
αB2
βB2
=
=
αA1
βA1
αA2
βA2
=
=
αC
βC
αO
βO
,
(14)
where δ21 = δi2/δi1 < 1 and δ32 = δi3/δi2 < 1.
Assuming that βi1 + γi1 ≫ αA1, βi2 + γi2 ≫ αA2 + βA1 , βi3 + γi3 ≫ βA2,
and the first forward and backward transitions are rate limiting ( βik ≫ δik
and γik ≫ αik, for k = 1 to 3) [18, 19], the occupation probabilities A1, A2
and A3 rapidly attain a quasi steady state
A1 ≈
A2 ≈
A3 ≈
,
,
αi1C1 + δi1B1
βi1 + γi1
αi2C2 + δi2B2
βi2 + γi2
αi3O + δi3B3
βi3 + γi3
,
(15)
(16)
(17)
and therefore, Eqs. (3) to (11) may be reduced to a six state master equation
(see Fig. 2)
dC1
dt
dC2
dt
dO
dt
= −(ρ1 + αC)C1(t) + βCC2(t) + σ1B1(t)
(18)
= αCC1(t) − (αO + βC + ρ2)C2(t) + βOO(t) + σ2B2(t)
(19)
= αOC2(t) − (βO + ρ3)O(t) + σ3B3(t)
(20)
5
dB1
dt
dB2
dt
dB3
dt
= ρ1C1(t) − (αB1 + σ1)B1(t) + βB1B2(t)
(21)
= ρ2C2(t) + αB1B1(t) − (αB2 + βB1 + σ2)B2(t) + βB2B3(t)(22)
= ρ3O(t) + αB2B2(t) − (βB2 + σ3)B3(t),
(23)
where the derived forward and backward rate functions for inactivation ρk
and σk are, in general, voltage dependent [20, 21, 22]
ρk ≈
σk ≈
αikγik
βik + γik
δikβik
βik + γik
,
.
(24)
(25)
If the conditions βik ≫ δik and γik ≫ αik for each k are not satisfied, the
inactivation of the Na+ current during a depolarizing potential may be bi-
exponential. From the assumptions of Eqs. (12) and (13), the inactivation
rate is not state dependent (ρk = ρ1 for each k), the deinactivation rates σ1 >
σ2 > σ2 ≈ 0, and therefore, from the microscopic reversibility conditions in
Eq. (14)
βB1αC < βC αB1,
βB2αO < βOαB2.
(26)
(27)
The nonzero eigenvalues of the characteristic equation for Eqs. (18) to
(23) are λj = −ωj for j = 1 to 5 where ωj may be approximated by the
roots ωkF and ωkG of the cubic polynomials F (ω) and G(ω) (see Fig. 3)
where ω1F < ω2F < ω3F , ω1G < ω2G < ω3G,
F (ω) = ω3 − a1ω2 + a2ω − a3,
(28)
a1 = αO + βO + αC + βC + ρ1 + ρ2 + ρ3
a2 = (βC + ρ2)(βO + ρ3) + ρ3αO +
(αC + ρ1)(αO + βO + βC + ρ2 + ρ3) − αCβC
a3 = αOρ3(αC + ρ1) + (βO + ρ3)(ρ2αC + ρ1βC + ρ1ρ2),
(29)
and, in this section, it is assumed that ρk = ρ1 for each k,
G(ω) = ω3 − b1ω2 + b2ω − b3
(30)
6
b1 = αB1 + βB1 + αB2 + βB2 + σ1
b2 = αB1(αB2 + βB2) + βB1βB2 + σ1(βB1 + αB2 + βB2)
b3 = σ1βB1βB2.
(31)
Therefore, for a depolarizing potential, we may define ωk ≈ ωkF , ωk+2 ≈
ωkG, for k = 2, 3 whereas for a hyperpolarizing potential, ωk ≈ ωkG, ωk+2 ≈
ωkF . If βh is the rate of inactivation and αh is the rate of recovery from
inactivation, it may be shown from the characteristic equation that ω1 =
αh + βh ≈ ω1G + ω1F where
ω1G =
ω1F =
b3
ω2Gω3G
a3
ω2F ω3F
,
.
(32)
(33)
If αi1, βi1, γi1 are, in general, exponential functions of V , the rate of
inactivation
βh ≈ ω1F = ρ1 =
αi1
1 + βi1/γi1
(34)
has an exponential voltage dependence for small clamp potentials but sat-
urates for a larger depolarization when αi1 is weakly dependent on voltage
(see Fig. 4) [1]. It is assumed that the activation sensors are independent
and hence αC = 2αm, αO = αm, βC = βm, βO = 2βm [2] where αm and
βm are HH rate functions for Na+ channel activation, and may be approx-
imated by two-stage expressions (see Fig. 4) [18]. If the DIII S4 sensor is
the slowest to deactivate (βB1 ≪ βB2) [9, 10], ω1G and ω2G are solutions of
the equation
ω2 − ω(αB1 + βB1 + σ1) + σ1βB1 = 0,
(35)
and the rate of recovery from inactivation αh ≈ ω1G. For the rate functions
of Fig. 4, αh ≈ σ1 when βB1 ≫ σ1 whereas for βB1 ≪ σ1 the rate of recov-
ery for inactivation αh ≈ βB1. From the microscopic reversibility conditions
of Eq. (14), we may assume that βB1 ∝ βC and αB1 ∝ αC and therefore,
αh(V ) and βm(V ) have a similar voltage dependence for small hyperpolar-
izing potentials, which is consistent with the HH determination of the rate
functions (βm(V ) ≈ 57αh(V )) [1].
If the Na+ channel is depolarized to a clamp potential V from a large
hyperpolarizing holding potential, the solution of Eqs. (18) to (23) for σ1 >
σ2 > σ3 ≈ 0 may be approximated by the solution of a master equation for
which σ1 > σ2, σ3 = 0
C1(t) = k1C1s + Σ5
j=1kj+1C1j exp(−ωjt)
(36)
7
where
(37)
(38)
(39)
(40)
(41)
(42)
C2(t) = k1C2s + Σ5
O(t) = k1Os + Σ5
B1(t) = k1B1s + Σ5
B2(t) = k1B2s + Σ5
B3(t) = k1B3s + Σ5
j=1kj+1C2j exp(−ωjt)
j=1kj+1Oj exp(−ωjt)
j=1kj+1B1j exp(−ωjt)
j=1kj+1B2j exp(−ωjt)
j=1kj+1B3j exp(−ωjt),
k−1
1
= Σ2
j=1(Cjs + Bjs) + Os + B3s,
C1s = σ1βB1βB2E0
C2s = σ1βB1βB2αC(βO + ρ1)
Os = σ1βB1βB2αCαO
B1s = βB1βB2a3
B2s = βB2(αB1 + σ1)a3 − ρ1σ1βB2E0
B3s = αB2(αB1 + σ1)a3 − ρ1σ1αB2E0 +
ρ1σ1αCαOβB1,
(43)
E0 = ρ1αO + (βO + ρ1)(βC + ρ1), a3 is defined in Eq. (29), the amplitudes
of the terms for each state are dependent on
C1j = E2(ωj)
C2j = −αC(ωj − βO − ρ1)
Oj = αC αO
B1j = −
B2j =
(−ρ1σ1E2(ωj) + (ωj − αB1 − σ1)F (ωj))
F (ωj)
σ1
1
σ1βB1
B3j = −
ρ1αCαO + αB2B2j
ωj − βB2
,
and
(44)
(45)
(46)
(47)
(48)
(49)
E2(ω) = ω2 − ω(αO + βO + βC + 2ρ1) + ρ1αO + (βO + ρ1)(βC + ρ1). (50)
Applying the initial conditions (C1(0) = 1 and C2(0) = O(0) = B1(0) =
B2(0) = B3(0) = 0), the parameters ki, i = 2 to 6 may be determined from
the solution in Eqs. (36) to (41). For a depolarizing potential, ω4 ≈ ω2G,
8
ω5 ≈ ω3G and therefore, from Eqs. (30) and (47) to (49) , assuming that
ω2 6= ω4 and ω3 6= ω5 for a coupled model of Na+ channel activation and
inactivation, F (ω2), F (ω3) ≪ F (ω4), F (ω5) and Bk2, Bk3 ≪ Bk4, Bk5
for each k. Therefore, to satisfy the initial conditions, k5, k6 ≈ 0 and
k2 =
k3 = −
k4 =
1 − k1σ1βB1βB2ω2ω3
(ω2 − ω1)(ω3 − ω1)
1 − k1σ1βB1βB2ω1ω3
(ω2 − ω1)(ω3 − ω2)
1 − k1σ1βB1βB2ω1ω2
(ω3 − ω1)(ω3 − ω2)
.
(51)
(52)
(53)
That is, each term of the open state probability in Eq. (38) with eigenvalue
λ = −ω4 or −ω5 where ω4 or ω5 may be approximated by the roots ω2G, ω3G
of the polynomial G(ω) has an amplitude close to zero. If it is assumed that
ρ1 = ρ2 = ρ3 and the two activation sensors are independent, the roots of
Eq. (28) are ω1F = ρ1, ω2F = αm + βm + ρ1, ω3F = 2(αm + βm) + ρ1 (see
Fig. 5) and
O(t) ≈ m(t)2h(t)
αm
m(t) =
h(t) =
(1 − exp[−(αm + βm)t])
αm + βm
αh + βh exp[−(αh + βh)t]
.
αh + βh
(54)
(55)
(56)
Therefore, following the application of a voltage clamp, the solution of
the master equation lies on an invariant manifold, defined by the eigenvec-
tors with eigenvalues that are determined by the roots of the polynomial
F (ω). If the deinactivation rates σ1 > σ2 > σ3 ≈ 0 are chosen to satisfy
microscopic reversibility, it may be shown from the numerical solution of
the master equation or from a more general form of the solution of Eqs.
(18) to (23) that O(t) = m(t)2h(t) is still a good approximation. From Eq.
(41), the probability for the inactivated state B3(t) has an initial delay that
diminishes with increasing depolarization, and may be approximated by a
bi-exponential function [6] (see Fig. 6).
Assuming that the time-dependence of the Na+ channel open state prob-
ability is described by the solution of a phenomenological master equation,
as well as the HH expression m(t)2h(t), the conditions for model reduc-
tion, F (ω2), F (ω3) ≪ F (ω4), F (ω5) for depolarizing potentials, provide con-
straints upon the choice of empirical activation rate functions. If ω2 ≈ ω4
and ω3 ≈ ω5 for a weakly coupled model of Na+ channel activation and
9
inactivation, these conditions are not satisfied and therefore, the terms with
eigenvalues −ω4 and −ω5 have a nonzero amplitude and also contribute to
the time-dependence of O(t).
When the Na+ channel is hyperpolarized to a clamp potential V from
a large depolarizing holding potential, the solution of Eqs. (18) to (23) for
σ1 > σ2, σ3 = 0 is given by Eqs. (36) to (41) where k1 and the stationary
solution are defined in Eqs. (42) and (43), and for j = 1 to 5
C1j =
σ1βB1βB2E2(ωj)
F (ωj)
C2j = −
σ1βB1βB2αC (ωj − βO)
F (ωj)
σ1βB1βB2αOαC
F (ωj)
Oj =
B1j = −βB1βB2
B2j = βB2 −
ρ1σ1E2(ωj)
F (ωj)
+ ωj − αB1 − σ1!
B3j = −
ρ1σ1αCαOβB1βB2 + αB2F (ωj)B2j
F (ωj)(ωj − βB2)
.
(57)
(58)
(59)
(60)
(61)
(62)
and E2(ω) is defined in Eq. (50).
For the nonzero eigenvalues, λj = −ωj for j = 1 to 5, of the characteristic
equation, ω1 = αh + βh ≈ ωkG + ωkF and we may define
ωk ≈ ωkG, ωk+2 ≈ ωkF
(63)
for k = 2, 3. Applying the initial conditions (C1(0) = C2(0) = O(0) =
B1(0) = B2(0) = 0 and B3(0) = 1), and assuming that ω2 6= ω4 and
ω3 6= ω5, from Eqs.
(28), (58) and (59), F (ω4), F (ω5) ≪ F (ω2), F (ω3),
C24 ≫ C22, C23, C25 and O4, O5 ≫ O2, O3. Therefore, to satisfy
the initial conditions, k5, k6 ≈ 0 and
k2 = −
k3 =
1
(ω2 − ω1)(ω3 − ω1)
1
(ω2 − ω1)(ω3 − ω2)
k4 = −
1
(ω3 − ω1)(ω3 − ω2)
.
(64)
(65)
(66)
Assuming that αC = 2αm, αO = αm, βC = βm, βO = 2βm, and that the
DIII S4 sensor is the slowest to deactivate (βB1 ≪ βB2) [9, 10], from Eq.
10
(36), we may write
C1(t) ≈ (cid:18)
X
βm
αm + βm(cid:19)2
(cid:18)1 − exp(−ω1t)(cid:20)1 +
ω1(1 − exp(−(ω2 − ω1)t)
ω2 − ω1
(cid:21)(cid:19) ,
(67)
where ω1 and ω2 are solutions of Eq. (35). Therefore, the time course of
the recovery from inactivation is bi-exponential and in agreement with the
kinetics determined from Nav1.4 channels [6] (see Fig. 7), but for large
negative potentials, ω2 ≈ βB1 ≫ ω1 ≈ σ1, and Eq.
(67) reduces to the
HH expression C1(t) = (βm/(αm + βm))2[1 − exp(−ω1t)]. For a weakly
coupled master equation, the conditions F (ω4), F (ω5) ≪ F (ω2), F (ω3) are
not satisfied and therefore, the terms with eigenvalues −ω4 and −ω5 also
contribute to C1(t).
When the Na+ channel conductance is regulated by the activation of
three voltage sensors in the DI, DII and DIII domains, and coupled to a two-
stage inactivation process where σ1 > 0 and σk = 0 for k > 1 (see Fig. 8), it
may be shown that during a depolarizing clamp potential, O(t) ≈ m(t)3h(t)
where m(t) and h(t) are defined in Eqs. (55) and (56), and αh and βh are
approximated by the the smallest roots of two quartic polynomials and may
be determined from Eqs. (34) and (35) (see Fig. 9). This description of the
time-dependence of the Na+ current is still a good approximation if it is
assumed that there is a separate opening step which is more rapid than the
activation of the S4 sensors [5]. Similarly, during a hyperpolarizing clamp
potential, if βB1 ≪ βB2, βB3 [9, 10], the time-dependence of the closed state
probability is a more general form of Eq. (67) (see Fig. 10)
C1(t) = (cid:18)
X
βm
αm + βm(cid:19)3
(cid:18)1 − exp(−ω1t)(cid:20)1 +
ω1(1 − exp(−(ω2 − ω1)t)
ω2 − ω1
(cid:21)(cid:19) .
(68)
MASTER EQUATION MODEL OF A Na+ CHANNEL WITH
A STATE DEPENDENT INACTIVATION RATE
In this section, we consider the effect of an increase in the inactiva-
tion rate as the S4 sensors activate (ρ1 < ρ2 < ρ3) [5, 6, 8], on the time-
dependence of m(t) and h(t). If it is assumed that αik = αi1, γik = γi1 for
each k and the DIV S4 rate functions satisfy
βi1 > βi2 > βi3,
(69)
11
the derived inactivation rate functions ρk are dependent on the closed or
open state. In order to satisfy microscopic reversibility, we may write
δ21
δ32
αB1
βB1
αB2
βB2
=
=
αA1
βA1
αA2
βA2
=
=
αC βi1
βC βi2
αOβi2
βOβi3
,
(70)
(71)
and therefore, from Eq. (13), the rate functions satisfy the inequalities (26)
and (27).
The eigenvalues of the characteristic equation may be approximated
by the roots of the cubic polynomials F (ω) and G(ω) (see Fig. 11), and
assuming that the activation sensors are independent, and ω1F = ∆1F ,
ω2F = αm + βm + ∆2F , ω3F = 2(αm + βm) + ∆3F are the roots of F (ω) in
Eq. (28), the rate of inactivation for a depolarizing potential is (see Fig. 12)
βh ≈
α2
mρ3 + 2αmβmρ2 + β2
mρ1
(αm + βm)2
(72)
which reduces to Eq.
(34) when the inactivation rate is not dependent
on the closed or open state. Therefore, the voltage dependence of βh has
contributions from the inactivation rate ρk for each k, as well as the ac-
tivation functions αm and βm. However, most of the voltage dependence
derives from the inactivation rate, and this is supported by the increase in
the time constant for inactivation in the charge-neutralized mutant Na+
channel DIV-CN [6].
For a hyperpolarizing potential, assuming that ω1G = ∆1G ≪ ω2G =
αB2 + βB1 + ∆2G ≪ ω3G = αB1 + βB2 + ∆3G are the roots of the polynomial
G(ω) in Eq. (30), it may be shown that the rate of recovery from inactivation
is
αh ≈
∆2G =
σ1βB1βB2
(αB2 + βB1 + ∆2G)(αB1 + βB2 − ∆2G)
D1 −qD2
2
1 − 4D2
,
(73)
(74)
where D1 = αB1 + βB2 − αB2 − βB1, D2 = αB1βB2 − αB1βB1 − αB2βB2 −
∆1G(αB1 + βB2) + σ1βB2. If αB1, ∆2G ≪ βB2, Eq. (73) reduces to
αh ≈
σ1βB1
αB2 + βB1 + ∆2G
,
(75)
12
and may be approximated by an exponential function of V when βB1 ≪
αB2 + ∆2G [1], whereas for more negative potentials, there is a gradual
increase of αh towards the saturation value σ1, in accord with the rate of
recovery for inactivated Na+ channels in hippocampal neurons (see Fig.
12) [5]. The deinactivation rate σ1 is only weakly voltage dependent for
hyperpolarizing potentials as βi1 ≫ γi1, and therefore, most of the voltage
dependence of αh derives from the activation and deactivation functions
between inactivated states. For the charge-neutralized mutant Na+ channel
DIV-CN, the voltage dependence of βi1(V ) is reduced so that βi1(V ) ≪ γi1
and σ1 ≪ δi1, but the voltage dependence of αB2 and βB1 are not affected,
and therefore, the expression αh is in accord with the data describing a slow
recovery from inactivation [6].
The solution of the master equation, Eqs. (18) to (23), for σ1 > σ2, σ3 =
0 and ρ1 < ρ2 < ρ3, is given by Eqs. (36) to (41) where the stationary
solution is
C1s = σ1βB1βB2E0
C2s = σ1βB1βB2αC(βO + ρ3)
Os = σ1βB1βB2αCαO
B1s = βB1βB2a3
B2s = βB2(αB1 + σ1)a3 − ρ1σ1βB2E0
B3s = αB2(αB1 + σ1)a3 − ρ1σ1αB2E0 +
ρ3σ1βB1αCαO,
(76)
E0 = E2(0) = ρ3αO + (βO + ρ3)(βC + ρ2), a3 is defined in Eq. (29), and the
amplitudes of the terms of each state are dependent on
C1j = E2(ωj)
C2j = −αC(ωj − βO − ρ3)
Oj = αC αO
B1j = −
F (ωj)
σ1
1
σ1βB1
B2j =
(−ρ1σ1E2(ωj) + (ωj − αB1 − σ1)F (ωj))
B3j = −
ρ3αCαO + αB2B2j
ωj − βB2
.
(77)
(78)
(79)
(80)
(81)
(82)
where E2(ω) = ω2 − (αO + βO + βC + ρ2 + ρ3)ω + (βC + ρ2)(βO + ρ3) + αOρ3.
13
From Eq. (38), applying the initial conditions (C1(0) = 1 and C2(0) =
O(0) = B1(0) = B2(0) = B3(0) = 0), it may be shown that the time-
dependence of the Na+ channel open state probability O(t) ≈ m(t)2h(t)
during depolarization to a clamp potential V (see Fig. 13), where h(t)
is defined in Eq.
(56), and the activation variable is dependent on both
activation and inactivation rate functions
αm
m(t) =
∆ =
(1 − exp[−(αm + βm + ∆)t])
αm + βm + ∆
αm(ρ2 + 2ρ3) + βm(2ρ1 + ρ2)
αm + βm
− 3βh.
(83)
(84)
The probability for entry into the inactivated state B3 may be also be ap-
proximated by a bi-exponential function (see Fig. 14), and the time course
of recovery from inactivation is given by Eq. (67) (see Fig. 15).
CONCLUSION
Hodgkin and Huxley described the time-dependence of the Na+ current
in the squid giant axon membrane in terms of the expression m(t)3h(t)
where the activation variable m(t) and inactivation variable h(t) satisfy rate
equations [1]. An alternative description of the Na+ current in nerve and
muscle membranes is provided by a master equation for coupled channel
activation and inactivation processes where the deinactivation rate to the
open state is small, but the rate to closed states increases as the activation
sensors in the domains DI, DII and especially DIII, deactivate. This model
accounts for the small Na+ current during repolarization of an inactivated
channel, the saturation of the rate of recovery from inactivation for large
hyperpolarized potentials and the delay in the time-course of the recovery
from inactivation [5]. If it is further assumed that inactivation is a two-stage
process, the model can account for the kinetics and voltage dependence of
Na+ inactivation for wild-type and mutant channels [6].
In this paper, we consider the coupling between two voltage sensors that
regulate the Na+ channel conductance and a two-stage inactivation process,
where the first forward and backward inactivation transitions of the DIV
S4 sensor are rate-limiting, ensuring that the inactivation decay during a
depolarizing voltage clamp is exponential. As the Na+ current following
inactivation is close to zero until the S4 sensors of the DIII, and either DI
or DII domains deactivate, we have assumed that σ1 > σ2 > σ3 ≈ 0, and
from the analytical solution of the reduced six state master equation for a
depolarizing clamp when the inactivation rate is uniform between states and
σ1 > σ2, σ3 = 0, the slowest eigenvalue is determined by the inactivation rate
ρ1, which has an exponential voltage dependence, but saturates for a large
14
depolarizing potential [1]. For a hyperpolarizing clamp of the Na+ channel,
the rate of recovery from inactivation is dependent on the deinactivation rate
σ1 to the first closed state, as well as the rate functions of the DIII S4 sensor
between inactivated states. The voltage dependence of the derived rate
functions for inactivation and recovery from inactivation have a similar form
to empirical expressions for Na+ channels in the squid axon [1], hippocampal
neurons [5] and Nav1.4 channels [6].
For a hyperpolarizing clamp potential, as the deinactivation rate σ1 >
σ2 > σ3 ≈ 0, it may be assumed that the deactivation rate functions be-
tween closed and open states are greater than those between inactivated
states (βO > βB2, βC > βB1), in order to satisfy microscopic reversibil-
ity. Therefore, the closed state terms with eigenvalues of the characteristic
equation that are determined by the roots of the polynomial F (ω) have an
amplitude that are close to zero, and as the DIII S4 sensor is the slowest to
deactivate (βB1 ≪ βB2) [9, 10], the time-dependence of the recovery from
inactivation is bi-exponential, and therefore, in agreement with the kinetic
data from Nav1.4 channels [6].
For a depolarizing clamp potential of a Na+ channel, assuming that
βO > βB2 and βC > βB1, each term of the open state probability with
eigenvalue λ = −ω where ω approximates a root of the polynomial G(ω),
also has an amplitude close to zero. A further simplification is possible when
it is assumed that the activation sensors are independent (αC = 2αm, αO =
αm, βC = βm, βO = 2βm) and it may be shown that the time-dependence of
the open state O(t) = m(t)2h(t). In most nerve membrane Na+ channels,
the activation of three voltage sensors regulate the Na+ channel conduc-
tance, and by application of similar constraints on the activation and deac-
tivation rate functions for inactivated and closed states, the time-dependence
m(t)3h(t) of the Na+ current may be derived from the solution to an eight
state master equation for coupled activation and inactivation. For models of
the Na+ channel where the inactivation rate from the closed or open states
increases as the S4 sensors activate, a more general form for the Hodgkin-
Huxley expression for the open state probability may be derived where m(t)
and h(t) are dependent on both activation and inactivation processes.
15
References
[1] A.L. Hodgkin and A.F. Huxley, J. Physiol. 117, 500 (1952).
[2] C. M. Armstrong, Physiol. Rev. 61, 644 (1981).
[3] B. Hille, Ion Channels of Excitable Membranes, 3rd ed. (Sinauer, Sun-
derland, M.A. 2001).
[4] J. Keener, J. Math. Biol. 58, 447 (2009).
[5] C-C. Kuo and B.P. Bean, Neuron 12, 819 (1994).
[6] D.L. Capes, M.P. Goldschen-Ohm, M. Arcisio-Miranda, F. Bezanilla
and B. Chanda, J. Gen. Physiol. 142, 101 (2013).
[7] B. Chanda and F. Bezanilla, J. Gen. Physiol. 120, 629 (2002).
[8] F. Bezanilla and C. M. Armstrong, J. Gen. Physiol.70, 567 (1977).
[9] A. Cha, P.C. Ruben, A.L. George, Jr., E. Fujimoto and F. Bezanilla,
Neuron 22, 73 (1999).
[10] M.F. Sheets, J.W. Kyle and D.A. Hanck, J. Gen. Physiol. 115, 609
(2000).
[11] P. G. DeCaen, V. Yarov-Yarovoy, E. M. Sharp, T. Scheuer and W. A.
Catterall, Proc. Natl. Acad. Sci. USA 106, 22498 (2009).
[12] R. Horn, S. Ding and H.J. Gruber, J. Gen. Physiol. 116, 461 (2000).
[13] C. M. Armstrong, PNAS 103, 17991 (2006).
[14] S.R. Vaccaro, J. Chem. Phys. 132, 145101 (2010).
[15] S.R. Vaccaro, J. Chem. Phys. 135, 095102 (2011).
[16] H. Lecar, H.P. Larrson and M. Grabe, Biophys. J. 85, 2854 (2003).
[17] H.A. Kramers, Physica. 7, 284 (1940).
[18] S.R. Vaccaro, Phys. Rev. E 90, 052713 (2014).
[19] J.L. Lacroix, F.V. Campos, L. Frezza, and F. Bezanilla, Neuron 79,
8651 (2013).
16
[20] M. Chahine, A.L. George Jr., M. Zhou, S. Ji, W. Sun, R.L. Barchi and
R. Horn, Neuron 12, 281 (1994).
[21] L. Q. Chen, V. Santarelli, R. Horn and R.G. Kallen, J. Gen. Physiol.
108, 549 (1996).
[22] M.F. Sheets and D.A. Hanck, J. Gen. Physiol. 106, 617 (1995).
17
C1
ΑC
ΒC
C2
ΑO
ΒO
O
Βi1
Αi1
Βi2
Αi2
Βi3
Αi3
A1
ΑA1
ΒA1
A2
ΑA2
ΒA2
A3
∆i1
Γi1
∆i2
Γi2
∆i3
Γi3
B1
ΑB1
ΒB1
B2
ΑB2
ΒB2
B3
Figure 1:
State diagram for Na+ channel gating where horizontal tran-
sitions represent the activation of two voltage sensors (DIII and either DI
or DII) that open the pore, and vertical transitions represent the two stage
inactivation process of the DIV voltage sensor and the inactivation motif.
18
C1
ΑC
ΒC
C2
ΑO
ΒO
O
Σ1
Ρ1
Σ2
Ρ2
Σ3
Ρ3
B1
ΑB1
ΒB1
B2
ΑB2
ΒB2
B3
Figure 2: The nine state system for Na+ channel gating in Fig. 1 may
be approximated by a six state system when βik ≫ δik and γik ≫ αik, for
k = 1 to 3, where ρk and σk are derived rate functions for a two-stage Na+
inactivation process, defined in Eqs. (24) and (25).
19
1
-
s
m
V
Ω
10
9
8
7
6
5
4
3
2
1
0
-150
-100
-50
V mV
0
50
Figure 3: The voltage dependence of ωj = −λj, j = 1 to 5, where λj is
a nonzero eigenvalue of the characteristic equation of the master equation
(solid line), and the voltage dependence of ωjF (dotted line) and ωjG (dashed
line), the roots of the cubic polynomials F (ω) and G(ω), where the rate
functions are αik(V ) = 1, γik(V ) = exp(3), βik(V ) = exp[−2(V − 3)/25]
for k = 1 to 3, δi1(V ) = 2.5, δi2(V ) = δi3(V ) = 0, αC = 2αm = αB1/3,
αO = αm = αB2/3, βC = βm = 83.3βB1, βO = 2βm = 10βB2 (ms−1) and
αm = 0.1(V + 25)/(1 − exp[−(V + 25)/10]) and βm = 4 exp[−(V + 50)/18]
are the HH rate functions for Na+ channel activation, assuming the resting
potential is V = −50 mV.
20
4
3
2
1
1
-
s
m
e
t
a
r
0
-150
-100
-50
V mV
0
50
Figure 4: Voltage dependence of the HH Na+ channel inactivation rate
functions βh = 1/(1 + exp(−(20 + V )/10)) and αh = 0.07 exp(−(V + 50)/20)
(dashed line) may be approximated by the analytical expression in Eq.
(34) and by the smaller root of Eq.
(35) (solid line), derived from a
master equation for a six state system where activation and two stage
inactivation are interdependent, and by the voltage dependence of ω1
determined numerically (dotted line) where the rate functions are de-
fined in Fig. 3. The HH Na+ channel activation rate functions αm =
0.1(V + 25)/(1 − exp[−(V + 25)/10]), βm = 4 exp[−(V + 50)/18] and
αm + βm (dot dashed line) may also be approximated by two stage ex-
pressions αm,2 = 2.3 exp[0.32(V + 50)/25](1 + 8.3 exp[−1.3(V + 50)/25]) and
βm,2 = 4.2 exp[−0.77(V + 50)/18](1 + 8.3 exp[−1.3(V + 50)/25]) (solid line)
[17].
21
t
O
1
0.8
0.6
0.4
0.2
0
t
2
C
t
1
C
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
V = 60 mV
V = -10 mV
V = -30 mV
0
1
2
3
4
5
6
V = 60 mV
V = -10 mV
V = -30 mV
0
1
2
3
4
5
6
V = 60 mV
V = -10 mV
V = -30 mV
0
1
2
3
4
5
6
t ms
Figure 5: During a depolarizing clamp potential for the six-state system in
Fig. 2 where activation and inactivation are interdependent, the open state
probability O(t) (solid line) ≈ m(t)2h(t) (dashed, dotted or dot-dashed),
C1(t) (solid line) ≈ [1 − m(t)]2h(t) (dashed, dotted or dot-dashed), C2(t)
(solid line) ≈ 2m(t)[1 − m(t)]h(t) (dashed, dotted or dot-dashed), where
m(t) and h(t) are solutions of rate equations for activation and inactivation,
and the rate functions are defined in Fig. 3.
22
t
3
B
1
0.8
0.6
0.4
0.2
0
V = 60 mV
V = -10 mV
V = -30 mV
0
1
2
3
4
5
6
t ms
Figure 6: During a depolarizing clamp potential for a six-state system, the
initial delay in the probability of the inactivated state B3(t) becomes less
pronounced as the clamp potential increases, and may be approximated by
a bi-exponential for V = -30 mV and 60 mV, and by a tri-exponential for V
= -10 mV (see Fig. 3).
t
1
C
1
0.8
0.6
0.4
0.2
0
V = -120 mV
V = -100 mV
V = -80 mV
0
1
2
3
4
5
6
7
8
t ms
Figure 7: During a hyperpolarizing clamp potential for a six-state sys-
tem, the first closed state probability C1(t) may be described by the bi-
exponential function in Eq. (67) (see Fig. 3).
23
C1
ΑC1
ΒC1
C2
ΑC2
ΒC2
C3
ΑO
ΒO
O
Σ1
Ρ1
Σ2
Ρ2
Σ3
Ρ3
Σ4
Ρ4
B1
ΑB1
ΒB1
B2
ΑB2
ΒB2
B3
ΑB3
ΒB3
B4
Figure 8: State diagram for Na+ channel gating where horizontal transi-
tions represent the activation of DI, DII and DIII voltage sensors that open
the pore, and vertical transitions represent the derived rate functions for a
two-stage Na+ inactivation process.
24
t
O
1
0.8
0.6
0.4
0.2
0
V = 60 mV
V = -10 mV
V = -20 mV
0
1
2
3
4
Figure 9: During a depolarizing clamp potential for the eight-state system
in Fig. 8, the open state probability O(t) (solid line) ≈ m(t)3h(t) (dashed,
dotted or dot-dashed), where m(t) and h(t) are solutions of rate equations for
activation and inactivation, and the rate functions are αik(V ) = 1, γik(V ) =
exp(3), βik(V ) = exp[−2.3(V − 13.2)/25] for k = 1 to 4, δi1(V ) = 2.5,
δik(V ) = 0 for k = 2 to 4, αm = 0.1(V + 25)/(1 − exp[−(V + 25)/10]),
βm = 4 exp[−(V + 50)/18], αC1 = 3αm, βC1 = βm, αC2 = 2αm, βC2 = 2βm,
αO = αm, βO = 3βm, αB1 = 3αC1, βB1 = 0.012βC1, αB2 = 3αC2, βB2 =
0.8βC2, αB3 = 3αO, βB3 = 0.8βO (ms−1).
25
t
1
C
1
0.8
0.6
0.4
0.2
0
V = -120 mV
V = -100 mV
V = -80 mV
0
2
4
6
8
10
12
t ms
Figure 10: During a hyperpolarizing clamp potential for an eight state
system, the first closed state probability C1(t) may be described by a bi-
exponential function in Eq. (68), where the rate functions are defined in
Fig. 9.
26
30
20
10
1
-
s
m
V
Ω
0
-150
-100
-50
V mV
0
50
Figure 11: The voltage dependence of ωj = −λj, j = 1 to 5, where λj is
a nonzero eigenvalue of the characteristic equation of the master equation
(solid line), and the voltage dependence of ωjF (dotted line) and ωjG (dashed
line), the roots of the cubic polynomials F (ω) and G(ω), where rate func-
tions are based on those determined for Nav1.4 channels [6], αik(V ) = 2.1,
γik(V ) = 24.9, for k = 1 to 3, βi1(V ) = 12.8 exp[−2.4V /25], βi2(V ) =
1.6 exp[−2.4V /25], βi3(V ) = 0.2 exp[−2.4V /25], δi1(V ) = 2.5, δi2(V ) =
δi3(V ) = 0, αC = 14.9 exp[0.3V /25] = αB1, αO = 7.45 exp[0.3V /25] =
αB2, βC = 0.8 exp[−0.9V /25] = 50βB1 and βO = 1.6 exp[−0.9V /25] =
3.3βB2(ms−1).
27
1
-
s
m
e
t
a
r
n
o
i
t
a
v
i
t
c
a
n
i
3
2
1
0
-200 -150 -100
-50
0
50
V mV
Figure 12: The voltage dependence of ω1, where λ = −ω1 is the slowest
eigenvalue (dotted line) of the solution of a master equation for a six state
system with state dependent inactivation, may be approximated by the rate
functions βh = 2.1/(1 + exp(−(V + 35)/10)) and αh = 0.065 exp(−(V +
80)/21) (dashed line) that have a similar voltage dependence to the HH
Na+ channel inactivation rate functions for the squid axon [1], and by the
expressions βh and αh in Eqs.
(72) and (75) (solid line) where the rate
functions are defined in Fig. 11.
28
t
O
1
0.8
0.6
0.4
0.2
0
t
2
C
t
1
C
1
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
0
V = 60 mV
V = -10 mV
V = -40 mV
0
1
2
3
4
5
6
V = 60 mV
V = -10 mV
V = -40 mV
0
1
2
3
4
5
6
V = 60 mV
V = -10 mV
V = -40 mV
0
1
2
3
4
5
6
t ms
Figure 13: During a depolarizing clamp potential for a six-state system with
state dependent inactivation, the open state probability O(t) (solid line)
≈ m(t)2h(t) (dashed, dotted or dot-dashed), the closed state probabilities
C1(t) (solid line) ≈ [1 − m(t)]2h(t) (dashed, dotted or dot-dashed), C2(t)
(solid line) ≈ 2m(t)[1 − m(t)]h(t) (dashed, dotted or dot-dashed), where the
rate functions are defined in Fig. 11.
29
t
3
B
1
0.8
0.6
0.4
0.2
0
V = 60 mV
V = -10 mV
V = -40 mV
0
1
2
3
4
5
6
t ms
Figure 14: During a depolarizing clamp potential for a six-state system
with state dependent inactivation, the initial delay in the probability of
the inactivated state B3(t) becomes less pronounced as the clamp potential
increases (see Fig. 11).
t
1
C
1
0.8
0.6
0.4
0.2
0
V = -160 mV
V = -120 mV
V = -100 mV
0
1
2
3
4
5
6
7
8
t ms
Figure 15: During a hyperpolarizing clamp potential for a six-state system
with state dependent inactivation, the first closed state probability C1(t)
may be described by the bi-exponential function in Eq. (67) (see Fig. 11).
30
|
1910.02901 | 1 | 1910 | 2019-10-07T16:43:31 | Combined force-torque spectroscopy of proteins by means of multiscale molecular simulation | [
"physics.bio-ph",
"cond-mat.soft"
] | Assessing the structural properties of large proteins is important to gain an understanding of their function in, e.g., biological systems or biomedical applications. We propose a method to examine the mechanical properties of proteins subject to applied forces by means of multiscale simulation. We consider both stretching and torsional forces, which can be applied independently of each other. We apply torsional forces to a coarse-grained continuum model of the antibody protein immunoglobulin G (IgG) using Fluctuating Finite Element Analysis and identify the area of strongest deformation. This region is essential to the torsional properties of the molecule as a whole, as it represents the softest, most deformable domain. We subject this part of the molecule to torques and stretching forces on an atomistic level, using molecular dynamics simulations, in order to investigate its torsional properties. We calculate the torsional resistance as a function of the rotation of the domain, while subjecting it to various stretching forces. We learn how these obtained torsion profiles evolve with increasing stretching force and show that they exhibit torsion stiffening, which is in qualitative agreement with experimental findings. We argue that combining the torsion profiles for various stretching forces effectively results in a combined force-torque spectroscopy analysis, which may serve as a mechanical signature for the examined molecule. | physics.bio-ph | physics |
Combined force-torque spectroscopy of proteins
by means of multiscale molecular simulation
Thijs W. G. van der Heijden1,*, Daniel J. Read2, Oliver G. Harlen2, Paul van der Schoot1,3, Sarah A. Harris4,5, and Cornelis
Storm1,6
1Theory of Polymers and Soft Matter, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The Netherlands
2School of Mathematics, University of Leeds, Leeds LS2 9JT, United Kingdom
3Instituut voor Theoretische Fysica, Universiteit Utrecht, Princetonplein 5, 3584 CC Utrecht, The Netherlands
4School of Physics and Astronomy, University of Leeds, Leeds LS2 9JT, United Kingdom
5Astbury Centre for Structural Molecular Biology, University of Leeds, Leeds LS2 9JT, United Kingdom
6Institute for Complex Molecular Systems, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, The
Netherlands
*Correspondence: [email protected]
ABSTRACT Assessing the structural properties of large proteins is important to gain an understanding of their function in,
e.g., biological systems or biomedical applications. We propose a method to examine the mechanical properties of proteins
subject to applied forces by means of multiscale simulation. We consider both stretching and torsional forces, which can be
applied independently of each other. We apply torsional forces to a coarse-grained continuum model of the antibody protein
immunoglobulin G (IgG) using Fluctuating Finite Element Analysis and identify the area of strongest deformation. This region
is essential to the torsional properties of the molecule as a whole, as it represents the softest, most deformable domain. We
subject this part of the molecule to torques and stretching forces on an atomistic level, using molecular dynamics simulations, in
order to investigate its torsional properties. We calculate the torsional resistance as a function of the rotation of the domain, while
subjecting it to various stretching forces. We learn how these obtained torsion profiles evolve with increasing stretching force and
show that they exhibit torsion stiffening, which is in qualitative agreement with experimental findings. We argue that combining
the torsion profiles for various stretching forces effectively results in a combined force-torque spectroscopy analysis, which may
serve as a mechanical signature for the examined molecule.
SIGNIFICANCE In this work, we propose a multiscale numerical approach to assess the mechanical properties of
macromolecules such as proteins. We perform a combined force-torque spectroscopy analysis on the mechanically most
relevant domain to compute the response signature of the spatial structure of the macromolecule. This information may
lead to a better understanding of molecular structure and function in biological context and may be used towards diagnostic
and sensing applications in the biomedical field.
INTRODUCTION
Proteins fulfil numerous different roles in organisms, such
as providing rigidity, transporting cargo through cells or
catalysing reactions. The functioning of a protein arises from
the folding of its intrinsic structure, a linear chain of amino
acids, into a higher-order hierarchical structure. Its final folded
structure consists of a certain shape with one or more active
sites, which facilitate the protein's function [1]. One type of
protein in particular, the antibody or immunoglobulin, plays
an important role in the mammalian immune system, by
specifically binding to foreign structures in the body. The fact
that its binding to a particular molecule is very specific makes
the antibody protein an excellent candidate to be employed
for analyte detection in a so-called immunoassay [2]. In an
immunoassay, the analyte is targeted by an antibody molecule
equipped with, e.g., a fluorescent or radioactive label. This
label can in turn be detected using conventional detection
methods.
In order to make immunoassays a viable method for point-
of-care diagnostics in medical applications, however, not only
the analysis, but also all of the sample preparation, transporta-
tion and mixing steps should be included in a 'lab-on-a-chip'
device. One proposed method for the integration is making
use of magnetic particles within the device [3, 4]. Not only
can these particles serve as labels for the analytes, but upon ac-
tuation with magnetic fields they can be actively manipulated
and be used to, e.g., mix fluids within the sensor. In addition,
the magnetic particles may act as magnetic tweezers in order
to exert forces on the molecules [4 -- 6]. This experimental
method is used to study the structural properties and unfold-
ing of molecules such as proteins, DNA and RNA [7 -- 11], and
may be employed side-by-side with atomic force microscopy
1
(AFM) [12 -- 16] and optical tweezers [17 -- 22] experiments to
assess the mechanical properties of molecules.
In recent years, van Reenen et al. [5] investigated the tor-
sional resistance of immunoglobulin protein complexes using
magnetic tweezers experiments. They were able to distinguish
between different torsion profiles for different proteins, a
quality that may eventually be employed to, e.g., identify the
nature of the binding (to be either specific or non-specific) in
immunoassays. To gain more insight into the relation between
the intrinsic structure of a molecule and its mechanical prop-
erties, we propose a multiscale numerical approach to analyse
the mechanical properties of proteins, in our case an im-
munoglobulin molecule, when subjected to externally applied
forces. Since investigating such large molecules (∼ 105 Da)
as a whole on an atomistic level is too costly in a computa-
tional context, we examine the molecule on a coarse-grained
mesoscopic level using Fluctuating Finite Element Analysis
(FFEA) [23 -- 25]. FFEA considers the overall shape of the
molecule and regards it as a continuum material internally.
This allows for a fast evaluation of the molecule subject to
thermal and/or external forces, at the expense of the loss of
information on the internal structure. Using FFEA, we can
identify the area in which immunoglobulin deforms the most
during torsion, which is presumably an essential region for the
molecule's torsional properties: it constitutes a weak link in
the rigidity of the molecule. Using molecular dynamics sim-
ulations, we perform a combined force-torque spectroscopy
analysis on this domain on a microscopic, full-atom level:
we investigate the torsional resistance of the molecule as a
function of the rotation, while a stretching force is exerted on
the structure. We extract the resulting torsion profile of the
molecule and learn how it develops as the stretching force
increases.
The remainder of this manuscript is organised as follows.
In the Methods section, we describe the subject molecule
of this work -- immunoglobulin G (IgG) -- in more detail.
We briefly describe the Fluctuating Finite Element Analysis
(FFEA) and molecular dynamics (MD) methods that we use
in our simulations. In the Results and Discussion section, we
present the results from the FFEA simulations, which we
use to define the domain of interest. We discuss our results
from the atomistic MD simulations of this relevant domain,
in which we subjected it to external forces and torques. In the
Conclusions section, we summarise our findings and draw
our conclusions.
METHODS
Force-torque spectroscopy on
Immunoglobulin G
Our subject molecule is an antibody protein, immunoglob-
ulin G (IgG), of which an atomistic structure was found by
X-ray diffraction (mouse IgG, Protein Data Bank: 1IGT) [26],
see Fig. 1 (top left). The protein consists of two structurally
2
identical heavy chains (red and yellow) and two identical light
chains (blue and green), with a total mass of ∼ 150 kDa. The
four chains combined form three bulky domains (the three
branches of the typical "Y"-shape), connected by a thin linker;
a feature shared by all isotypes of immunoglobulin [27].
Figure 1: The overall shape of the atomistic structure (top
left) of IgG is converted into a volume mesh (bottom right).
Schematic (top right): the torsional force is exerted on the
molecule at the bottom (red shaded area), while the protein is
fixed at the ends of the top branches (yellow shaded areas), to
mimic the fixation to a substrate. The torsion axis and rotation
direction are indicated.
The IgG molecule was the subject of experimental work
by van Reenen et al. [5], who investigated the torsional proper-
ties of a protein complex, formed by either two IgG molecules
or an IgG and a protein G molecule. They sandwiched the
complex between a glass substrate and a magnetic particle and
found that the different protein complexes respond differently
to exerted torques and that they stiffen for increasing torsion
angles: they exhibit a torsion stiffening behaviour. We note
that in practice, in addition to the torsional forces, a stretching
force may be exerted on a protein complex by applying a
second magnetic field that pulls the magnetic particle away
from the substrate. However, stretching forces are already
inherently present in the experimental situation due to the
gravitational forces that work on the magnetic particle and the
forces that arise from the direct interaction between the mag-
netic particle and the substrate to which the protein complex
is bound.
Considering the addition of such stretching forces in
the torsion experiments allows for a second, independent
axis to exert forces on the molecules on. Arguably, both the
conventional stretching and torsional rigidities of a molecule
depend in fact simultaneously on both the amount of stretching
and the amount of torsion. This creates a direct coupling
between the stretching and the torsion in such a combined
force-torque spectroscopy analysis, and may result in a more
complex mechanical "signature" of a certain structure. We
aim to numerically investigate the mechanical properties of a
single IgG molecule, and to calculate the torsion profile of
part of IgG under the influence of stretching forces. To that
end, we first identify our region of interest on a mesoscopic
level using Fluctuating Finite Element Analysis (FFEA). We
briefly discuss the method below.
Mesoscopic simulation using
Fluctuating Finite Element Analysis (FFEA)
The FFEA method treats large molecules such as proteins as
a continuum material, in order to simulate their behaviour
on a mesoscopic scale [23, 28]. The underlying principle
is that the overall shape of such a molecule determines its
function, and that the intrinsic structure of the molecule is
of less importance. We presume that the dynamics of such a
material is described by the Cauchy momentum equation,
+ (uuu · ∇∇∇)uuu =
∂uuu
∂t
∇∇∇ · σσσ,
1
ρ
(1)
where uuu denotes the velocity at any point in the material, ρ
is the mass density of the material and σσσ is the total stress
exerted by the continuum material. We employ the Kelvin-
Voigt material model, which allows us to write the total stress
σσσ as the sum of elastic, viscous and thermal stress terms,
σσσ = σσσe + σσσv + σσσt.
(2)
The elastic stress σσσe is based on a Mooney-Rivlin hyperelastic
model for the stress [29, 30], which is described in more
detail in the Appendix. For the expressions for the viscous
and thermal stresses we refer the reader to Ref. [23], since
these are of secondary importance in this study. The protein's
material properties are parametrised using continuum material
parameters, such as the mass density ρ, the bulk modulus K
and the shear modulus G. To directly measure experimentally
the values for these parameters is not straightforward, although
estimations can be made by considering experimental data
on the density of proteins (roughly 1.5 g/cm3) [31, 32], their
internal viscosity (∼ 1 mPa s) [33] and their elastic modulus
(order 107 − 108 Pa) [33, 34]. We note, however, that the
location of the most deformed area is arguably more sensitive
to the geometry of the molecule than to the exact values for
the material properties, since we exert external forces on the
molecule and parametrise the structure homogeneously.
We create the volume mesh for the molecule by consider-
ing the atomistic model, converting it to an electron density
map and meshing its surface [28], as illustrated in Fig. 1.
We coarsen the surface until the shortest edge is 7 Å, while
maintaining the volume [33], and create the volume mesh
using the Netgen software package. We investigate the IgG
molecule in numerical torsion simulations by exerting an ex-
ternal torque on the bottom branch of the molecule, see Fig. 1.
In order to mimic the fixation to a substrate, we immobilise
the ends of the top two branches. We note that we do not exert
stretching forces on the molecule in these FFEA simulations.
In order to avoid strong sudden deformations of the molecule,
we slowly increase the total torque on the molecule up to
various constant values τ. We exert the torque by adding an
additional torsion force to all the mesh nodes within the red
shaded area in Fig. 1, where the magnitude depends on the
distance to the torsion axis. For definiteness, we disregard
the thermal stresses in the material, and rather focus on the
viscoelastic response of the structure to the external torque.
To that end, we set the thermal stresses in our simulations to
zero.
By studying the internal stresses in the molecule during
torsion using FFEA, we find the domain of interest for this
protein. We isolate the domain and analyse it using molecular
dynamics simulations. We briefly describe the simulations
below.
Atomistic simulation using
molecular dynamics (MD)
We perform molecular dynamics (MD) simulations on the
relevant domain of the molecule using the GROMACS soft-
ware package [35]. The simulation is performed in an implicit
water solution, with a 100 mM monovalent salt concentration,
using a Langevin thermostat with friction constant γ = 5 ps−1,
in order to ensure a strongly damped dynamics. This is pru-
dent in order to suppress the influence of the high rotation
rate (as discussed below) and to minimise inertial effects
in our analysis of the mechanical response. We employ the
Amber ff99SB-ILDN forcefield for the parametrisation of the
interactions in the atomistic representation of the protein [36].
By simultaneously exerting a stretching force f and a
torque τ on the structure we are able to explore its mechanical
properties in this combined force-torque spectroscopy analysis.
We exert various stretching forces f on the molecule, ranging
from 0 to 3200 kJ mol−1 nm−1 (approximately 5.3 nN). We
exert the torque τ on the molecule by rotating a harmonic
potential well V around the central axis of the molecule,
which the forced residues of the structure are pulled into
(GROMACS: Vrm2) [37], see Fig. 2. The potential minimum
is indicated by the dashed line, and the blue and red colours
represent low and high potential energy, respectively. The
rotation rate is 100 ° ns−1 and the magnitude of the torque
depends on the positions of the residues as well as the spring
constant k associated with the potential well. k ranges from
0 to 3000 kJ mol−1 nm−2 (∼ 5 N m−1) in our simulations,
resulting in torques τ up to approximately 2000 kJ mol−1
(∼ 3300 pN nm). For reference, in the experiments by van
Reenen et al. [5], the maximum exerted torques are of the
order of 4000 pN nm.
In the next section, we present our results on the meso-
scopic FFEA simulations in order to identify the domain of
interest in the molecule. We isolate this area and perform
3
Figure 2: Schematic of the potential well V, that rotates in
a counter-clockwise direction with a rate of 100 ° ns−1. Blue
indicates low potential energy, red indicates high potential
energy. The potential minimum is indicated by the dashed
line. The forced residue, indicated by the blue circle, is pulled
towards the bottom of the well, resulting in the torque τ. The
central axis (the gray dashed line in Fig. 4) is indicated by (cid:12).
full-atom MD simulations on it, subjecting it to external forces
and torques. We analyse the torque as a function of the torsion
of the molecule and the exerted stretching force, resulting in a
combined force-torque spectroscopy analysis of the domain.
RESULTS AND DISCUSSION
Mesoscopic FFEA simulations
We perform mesoscopic simulations of the IgG molecule
subject to torsional forces using FFEA. In Fig. 3 we show
snapshots of a typical torsion simulation for various values of
the torsion angle φ. We measure the stress at each position
in the molecule and shade the stressed regions in red. The
intensity of the red colour indicates the amount of stress,
which is normalised to the maximum stress in each snapshot.
We find that for increasing torsion angles φ, the stress
in the linker area between the three bulky domains strongly
increases. The high stress indicates a strong deformation in this
area, which hints at this region being critical for the torsional
resistance of the molecule as a whole: it constitutes the softest
area in the structure. We note that in this particular case this
is not entirely surprising, considering that it is a relatively
thin linker in the otherwise bulky geometry of the molecule.
However, we argue that in general this need not be the case.
The use of the FFEA continuum model to assess the magnitude
of the stress as a function of position within a molecule enables
us to select which region is necessary to consider for more
detailed analysis using atomistic simulations, which account
for the internal structure of the protein [38].
We further investigate the linker domain and its mechanical
properties on a microscopic level: we isolate the linker region
and subject it to torques and stretching forces in molecular
dynamics simulations in order to extract its torsion profile for
various stretching forces. We discuss the results below.
4
Figure 3: Snapshots of a typical FFEA torsion simulation
of the IgG molecule for various torsion angles φ, for a total
torque τ = 299 pN nm. The surface elements shaded in red
indicate the areas in which the total stress is relatively high.
The intensity of the red shade is scaled to the maximum stress
magnitude σmax in each snapshot. The areas of high stress
indicate strong deformations.
Atomistic MD simulations
The flexible linker region of the IgG molecule consists of two
identical peptide chains with 13 residues each, interconnected
by three disulfide bonds, see Fig. 4. Note that only two of the
disulfide bonds (in yellow) are clearly visible.
We first perform an energy minimisation and a NVT
simulation, while keeping the ends of the chains immobilised,
to equilibrate the molecule. Subsequently, we perform a
molecular dynamics simulation while subjecting the molecule
to a stretching force f along its central axis, in the absence
of a torque. We fix the top residues (indicated in red) in
place, to mimic the fact that in reality they are connected to
the rather bulky top two branches of the molecule. For the
initial stretching, we allow the bottom residues (indicated in
blue) to only move strictly in the direction of the stretching
force, as this serves as an equilibration for the length of the
chains. Then, the bottom residues are released and a torque τ
is exerted on them along the molecule's central axis, which
causes the structure to rotate.
Directly calculating the rotation angle φ of the bottom
residues around the torsion axis is only sensible if the molecule
remains reasonably stretched during the torsion. Due to the
double-chain nature of the structure, however, exerting a strong
torque on the molecule not only results in a rotation around
the torsion axis, but it also causes additional coiling of the
central axis of the molecule. This coiling is not accounted for
if we directly measure φ. In order to capture all of the torsion
in the structure, we consider the twist Tw of the backbones of
Figure 4: The linker region, consisting of two identical protein
chains (indicated by the red and blue ribbons), in a full-atom
representation. The central axis of the molecule is indicated
(gray dashed line), as well as the direction of the torque τ and
the pulling forces f . The residues at the top (red circles) are
fixed to their position, the torque and force are exerted on the
residues at the bottom (blue circles).
the chains in the linker region as a measure for the amount of
rotation contained in the molecule. The twist is a quantity of
a mathematical ribbon, independent of an external reference
axis, that describes the winding of the ribbon around itself
with respect to the ribbon axis [39]. We note that the value
for Tw becomes negative if we move along the ribbon in the
opposite direction. We construct the ribbon by considering the
pairs of corresponding atom between the backbones of the two
identical chains, from the lowest disulfide bond, up to the top
residues. We define the ribbon axis as the average positions
of each pair of atoms and the ribbon boundary consists of one
of the two peptide chains. The protocol to calculate the twist,
as well as the notion that using the writhe as an alternative
measure of the torsion of the molecule turns out to not be
viable, are discussed in detail in the Appendix.
In our torsion simulations, we calculate the twist Tw
of the structure for each snapshot of the simulation and
track the exerted torque τ. Figure 5 shows the data from a
typical simulation, for which we arbitrarily set the stretching
force f = 800 kJ mol−1 nm−1 and the spring constant k =
50 kJ mol−1 nm−2, which results in a regularly oscillating
behaviour. We show the data for the twist Tw in blue, and
the torque τ in red. The light colour represents the original
data, the darker colour shows a weighted running average
of the data over time, in order to highlight the trends. The
weighted running average ¯xi at a point in time i is calculated
by taking 2N data points surrounding i, and weighing them
Figure 5: The twist Tw (blue) and applied torque τ (red) as a
function of time, for a stretching force f = 800 kJ mol−1 nm−1
and a spring constant k = 50 kJ mol−1 nm−2. The light colour
represents the original data, the darker colour shows a weighted
running average of the data over time, calculated using Eq. (3).
by the inverse distance to the point,
N
N
¯xi =
j=−N xi+j/( j + 1)
1/( j + 1)
j=−N
,
(3)
where xi is the original data at the point in time i. For the
trends in Fig. 5, we set N = 30.
We see that the data for both the twist and the torque
strongly fluctuate as a result of the Brownian motion of
the molecule, but that overall the twist and torque show
coherent behaviour. At very short times (t < 100 ps), we see
a fairly constant value for Tw, and no torque τ is exerted.
This corresponds to the length equilibration step, where we
stretch the molecule using a force f without exerting a torque.
Subsequently, we release the bottom residues and rotate the
potential well V. As a result, we see an increase of both Tw
and τ in time. After a while, the molecule reaches a state of
oscillatory motion, as both Tw and τ reach a maximum value,
after which the molecule partially relaxes to a less strained
state. The process repeats for each (half-)cycle of the rotating
potential well: as long as the rotated residues remain near
the potential minimum, hardly any torque is exerted. Upon
further rotation, however, the torsional resistance increases
and a torque is exerted on the residues in order to enforce
the rotation. As the torsional resistance of the molecule then
becomes too strong for the potential well to overcome, the
residues move out of the well and the structure is allowed
to partially relax until the rotating potential catches up to it
again. Since the potential well is symmetric with respect to
the central axis (as is schematically shown in Fig. 2), this
results in a period for the oscillatory motion of half a rotation
of the well, corresponding to 1.8 ns for a rate of 100 ° ns−1.
We repeat the torsion simulations for various values of the
spring constant k for the rotating potential well V. We track the
5
CONSH01020304050time[ns]−1.5−1.0−0.50.00.51.01.5twist[−]−50050100150200torque(cid:2)kJmol−1(cid:3)twist Tw and the torque τ during the simulations and combine
the results into a torsion profile for a given stretching force f :
we show the exerted torque τ as a function of the twist Tw of
the molecule. In Fig. 6 we show the torsion profiles for various
values of the stretching force f : (a) 0, (b) 200, (c) 800, and
(d) 3200 kJ mol−1 nm−1. All individual data points are shown
with a 5% opacity, which effectively results in a configuration
density plot of the molecule in torque-twist space. The darker
lines are trends, which we calculate by averaging the torques
τ for twists Tw between −1.5 and 1.5, in steps of ∆Tw = 0.02.
The error bars shown are the standard deviations around the
averages of the torque. Figure 6e shows typical snapshots of
the simulations for specific values of the twist Tw = −1, 0, 1,
for f = 3200 kJ mol−1 nm−1.
We learn from Fig. 6 that regardless of the magnitude of
the stretching force f , the molecule initially exhibits a torsion
stiffening behaviour: the torque τ increases non-linearly with
increasing Tw. For Tw (cid:46) 0.1 we see that the profile remains
fairly flat and close to τ ≈ 0, which hints at a low torsional
resistance, whereas for increasing values the profile steepens,
indicating an increase of the torsional stiffness: the torsional
resistance (or "torsional spring constant") is associated with
the derivative of the torque with respect to the twist. The
stiffness increase is in agreement with the results of van
Reenen et al. [5], who reported torsion stiffening behaviour
of the IgG-IgG complex for increasing torsion angles (up
to approximately 250°). In our analysis, however, for even
greater rotations the torsion profile appears to flatten off. We
return to this phenomenon below.
We note that for f = 0, in Fig. 6a, the spread of the data
points is notably larger than for nonzero stretching forces. This
is caused by the fact that the molecule may become strongly
supercoiled, and the configuration of the molecule regularly
becomes inverted (i.e., the bottom residues end up above the
top residues) due to the strong internal stresses caused by
the exerted torque and the absence of a stretching force. As
a result, the torque needed to reach a certain twist varies
rather strongly. Nevertheless, these data points do display
characteristics for the mechanical properties of the molecule
at zero stretching force. The profile exhibits a mild torsion
stiffening and is fairly symmetric for positive and negative
values of Tw.
As we increase the stretching force f , in Figs. 6b and c,
the standard deviation decreases and the underlying shape of
the torsion profile becomes more apparent: the two chains
are pulled towards each other as a result of the stretching
and cause the onset of an internal counterforce. The shapes
of the two torsion profiles are rather similar, the profile for
f = 800 kJ mol−1 nm−1 (Fig. 6c) displaying a slightly stiffer
profile than the profile for f = 200 kJ mol−1 nm−1 (Fig. 6b),
as may be expected. For both profiles, we see that in the flat
regime, for small twists (Tw (cid:46) 0.1), the spread is small,
compared to the data for stronger twists. This indicates that
the torsional stiffness of the molecule is rather insensitive to
its exact configuration in the untwisted state. If the molecule
6
Figure 6: The development of the twist-torque profile for
increasing stretching force; the exerted torque τ is shown as a
function of the twist Tw for various stretching forces f : (a) 0,
(b) 200, (c) 800, and (d) 3200 kJ mol−1 nm−1. (e) Snapshots
of the torsion simulation illustrating the conformation of the
molecule at specific values of the twist Tw = −1, 0, 1, for
f = 3200 kJ mol−1 nm−1.
is twisted more strongly, its torsional response depends more
strongly on how the chains are positioned with respect to
each other. While the torsion profiles appear rather symmetric,
some features arise that may be associated with the internal
structure of the molecule. For example, the magnitude of
the torque is greater for negative values of the twist than
for positive ones. This hints at a preferred rotation direction
for the molecule, which may indicate that the helicity of the
structure is slightly right-handed. In addition, we find that
"islands" form in torque-twist space, around Tw = ±1, and
that their shape depends on the sign of the twist Tw and the
magnitude of the stretching force f .
If we stretch the molecule strongly, with a force f =
3200 kJ mol−1 nm−1, shown in Fig. 6d, we find that the features
become more pronounced and asymmetric. Still, the profile is
rather flat for small values of Tw, however, the aforementioned
islands now have a distinctly different shape for positive or
negative twists. This indicates that for strong stretching forces
f , the details in the structure of the molecule may become
crucial for the torsional resistance of the protein, and that as a
result the asymmetries in the torsion profile arise.
If we join together the torsion profiles for different stretch-
ing forces f , we practically create a combined force-torque
spectroscopy analysis of the structure, see Fig. 7. We only
Figure 7: The trends of the combined force-torque spec-
troscopy profile of the IgG linker region: the exerted torque τ
is shown as a function of the twist Tw, for various values of
the stretching force f . For clarity the individual data points
and the standard deviations of the torques are not shown.
show the trends without the standard deviations, in order to
illustrate the overall evolution of the torsional stiffness profile.
Note that the full shape of the profile does contain more infor-
mation about the structural details and the resulting torsional
behaviour, and that the shown trends may not necessarily fully
represent the shape of the torsion profile, which is apparent
from the trendline in Fig. 6d. From Fig. 7 we learn that the
molecule indeed gradually stiffens as the stretching force f
increases. It exhibits a torsion stiffening behaviour for small
twists, after which the resistance flattens off and the torsional
resistance remains approximately constant. Given the double-
chain structure of the molecule, we in fact expect the profiles to
show an additional torsion stiffening for even greater values of
Tw (as the structure becomes more strongly coiled), however
we have not investigated greater rotations. We speculate that
we do see the onset of such additional stiffening for strongly
negative twists in the curves for f = 1600 kJ mol−1 nm−1 and
f = 3200 kJ mol−1 nm−1. Figure 7 indicates that the end of
the torsion stiffening regime, represented by the inflection
point of the torque-twist curve, gradually shifts to smaller
values of Tw as f increases. In addition, the asymmetry
between the magnitude of the exerted torques for positive and
negative twists is apparent.
We argue that such an evolution of the shape of the
torsion profile of a molecule with increasing stretching force
composes a mechanical signature for said molecule. It may
facilitate the development of a characterisation method for
proteins or other large molecules, for which we can predict
the mechanical response by performing this numerical force-
torque spectroscopy analysis.
This concludes the discussion of our results. In the next
section, we summarise our main findings and draw our con-
clusions.
CONCLUSION
We put forward a multiscale molecular simulation method
to perform a combined force-torque spectroscopy analysis of
large molecules such as proteins: we analyse the mechanical
response of a molecule when subjected to external torques
and stretching forces. We combine (1) Fluctuating Finite El-
ement Analysis (FFEA) with (2) molecular dynamics (MD)
simulations that incorporate external forces. We find that
using FFEA, we are able to indicate the region within an IgG
molecule that is crucial for its torsional rigidity: while subject-
ing the molecule to torques, we locate the area of strongest
deformation, suggesting that this linker domain is likely to be
the most flexible. We subsequently isolate the relevant domain
and investigate its torsional properties using MD simulations,
while subjecting it to stretching forces and torques. We find
that the linker region exhibits a torsion stiffening behaviour,
a result that is in qualitative agreement with experimental
results by van Reenen et al. [5]. For stronger rotations, the
exerted torque flattens off and the torsional resistance remains
approximately constant. As we increase the stretching force
exerted on the molecule, we find that the structure stiffens
and that features and asymmetries arise in the shape of the
torsion profile, which indicates that the structural details of
the molecule may be crucial for its mechanical response.
Combining the torsion profiles for different stretching forces
effectively results in a combined force-torque spectroscopy
analysis of the molecule and we argue that this may be used
7
−1.5−1.0−0.50.00.51.01.5twist[−]−1000−50005001000torque(cid:2)kJmol−1(cid:3)force(cid:2)kJmol−1nm−1(cid:3)020040080016003200−540−360−1800180360540twistangle[◦]as a characterisation method for the examined structure.
In conclusion, our study serves as a proof of concept for
an efficient numerical evaluation of the mechanical response
of a large molecule. This method facilitates the automation of
the multiscale procedure for a high-throughput computational
analysis of multiple proteins subject to stretching and torsional
forces. If the atomistic structure of such a molecule is known,
we may disregard the domains less relevant to the torsional
stiffness and focus on the softest areas from the perspective
of the torsional rigidity, aided by the FFEA method. Bulky
domains within the molecule that consist of many atoms and
are likely to be relatively rigid, do not need to be taken into
account. This is a considerable advantage for the atomistic
molecular dynamics simulation that is subsequently used to
investigate the torsional properties of the relevant domain in
more detail.
AUTHOR CONTRIBUTIONS
TWGvdH performed simulations, analysed the results and
wrote the manuscript. DJR, OGH and SAH developed the
FFEA simulation method. TWGvdH, PvdS, SAH and CS
designed the research and interpreted the results. All authors
contributed to the manuscript.
ACKNOWLEDGMENTS
Molecular images were produced using the VMD software
package [40]. The authors thank Ben Hanson, Fabiola Gutiérrez-
Mejía and René de Bruijn for fruitful discussions. TWGvdH
thanks the University of Leeds, where part of this work was
executed, for their hospitality.
APPENDIX
The elastic stress
The elastic energy of the molecule is based on a Mooney-
Rivlin hyperelastic model for the stress [29], with an adaptation
proposed by Gent [30], in order to introduce a maximum de-
formation for the structure. The strain energy density function
W is as follows:
W = − G
+
1 − I1 − 3
Im − 3
(cid:18)
(cid:19)
J2 − 1(cid:17) − 3K + 4G
FFFFFFT(cid:17) ;
6
12
(cid:16)
2 (Im − 3) ln
3K − 2G
I1 = Tr(cid:16)
J = Det(FFF) = V/V0,
ln J,
(4)
(5)
(6)
with G and K the shear and bulk moduli, respectively, and I1
and J two invariants of the deformation gradient tensor FFF:
with V and V0 the instantaneous and initial volumes, respec-
tively. Im is the maximum value for I1 for the structure. In our
8
analysis, in order to strongly limit the deformations and to em-
phasise the stressed regions, we set Im = 3.1. FFF describes the
deformation for a mapping of a position XXX to a new position
xxx: XXX (cid:55)→ xxx(XXX):
Fi j =
∂xi
∂Xj
.
(7)
The elastic stress tensor is derived from the strain energy
density function as follows:
σσσe =
FFFT ,
1
J
∂W
∂FFF
(cid:18) 3K − 2G
6J
(cid:16)
(cid:19)
J2 − 1(cid:17) − G
J
(8)
III,
(9)
which results in:
Im − 3
Im − I1
σσσe =
G
J
FFFFFFT +
with III the identity matrix.
Calculating the twist
The twist of a ribbon indicates the amount of (right-handed)
winding of the ribbon around itself, along the central axis. We
can define a ribbon by considering the central axis C1 and the
boundary C2, see Fig. 8.
C1
C2
nnn
ttt
Figure 8: A ribbon with central axis curve C1 and boundary
curve C2. The unit tangent to the central axis ttt and the unit
normal vector nnn pointing from C1 to C2 are indicated. This
particular ribbon contains a twist Tw = 1.
We indicate the unit tangent to the central axis ttt(s) and
the normal unit vector nnn(s) perpendicular to ttt, pointing from
C1 to C2. They share a common arc length parameter s. The
twist Tw of the ribbon can be calculated as [39],
Tw =
1
2π
(ttt × nnn) · dnnn
ds
ds.
(10)
C1
In order to calculate the twist contained in the linker region
(see Fig. 4), we first need to transform this region to a ribbon
representation. We consider only the backbones (consisting
of carbon and nitrogen atoms) from the lowest disulfide bond
up to the top. These are the atoms corresponding to the green
ribbons in Fig. 6e. Taking into account the dangling ends at
the bottom of the two chains in the twist calculation would
result in overestimations of the intrinsic twist of the protein,
since these are hindered less by the two-chain structure. As the
two chains are structurally identical, it is prudent to connect
the corresponding atoms in order to form a ribbon. We define
Þ
the central axis AAAi as the average positions of the connected
atoms in each chain,
AAAi =
rrr1,i + rrr2,i
2
,
(11)
where rrr j,i denotes the position of the i-th atom in chain j. The
discrete ribbon is now defined by the central axis AAAi and the
boundary BBBi ≡ rrr1,i, see Fig. 9.
AAAi
BBBi
BBBi+1
AAAi+1
Figure 9: A discrete ribbon with central axis positions AAAi and
boundary positions BBBi.
We calculate the twist Tw of the discrete ribbon as
n−1
i=1
Tw =
1
2π
where
αi arccos(vvvi · wwwi) ,
αi = sgn[AAAiAAAi+1 · (vvvi × wwwi)] ;
vvvi =
BBBiAAAi × BBBiAAAi+1
BBBiAAAi × BBBiAAAi+1 ;
BBBi+1AAAi × BBBi+1AAAi+1
BBBi+1AAAi × BBBi+1AAAi+1 ,
wwwi =
(12)
(13)
(14)
(15)
and AAAiAAAi+1 indicating the segment from position AAAi to AAAi+1 [39].
The factor αi accounts for the direction of the twist, i.e.,
whether the respective segments cause a positive or negative
contribution to the total twist.
Considering the writhe
In our analysis of the rotation of the molecule we also con-
sidered using the writhe of the structure. In order to calculate
the writhe, we construct a closed loop consisting of the two
backbones of the chains, the lowest disulfide bond connecting
the two chains and an artificial connection between the top
two residues. The writhe is a quantity of a closed loop, which
means that it is independent of any external reference axis
or orientation. However, it turns out that the writhe is rather
sensitive to fluctuations in the positions of the atoms. This
results in uncertainties and inaccuracies in determining the
rotation of the molecule and we therefore deem it unsuitable
for our analysis.
REFERENCES
[1] Bourne, P. E., and H. Weissig, 2003. Structural Bioin-
formatics. Wiley-Liss, Inc., Hoboken, New Jersey.
[2] John, R., C. Sheehan, S. Binder, and J. He, 2013. The
Immunoassay Handbook - Theory and Applications
of Ligand Binding, ELISA and Related Techniques.
Elsevier.
[3] van Reenen, A., A. M. de Jong, J. M. J. den Toonder, and
M. W. J. Prins, 2014. Integrated lab-on-chip biosens-
ing systems based on magnetic particle actuation -- a
comprehensive review. Lab Chip 14:1966 -- 1986.
[4] Moerland, C. P., L. J. van IJzendoorn, and M. W. J.
Prins, 2019. Rotating magnetic particles for lab-on-
chip applications -- a comprehensive review. Lab Chip
19:919 -- 933.
[5] van Reenen, A., F. Gutiérrez-Mejía, L. van IJzendoorn,
and M. Prins, 2013. Torsion Profiling of Proteins Using
Magnetic Particles. Biophys. J. 104:1073 -- 1080.
[6] Gutiérrez-Mejía, F., L. van IJzendoorn, and M. Prins,
2015. Surfactants modify the torsion properties of
proteins: a single molecule study. N. Biotechnol. 32:441 --
449.
[7] Smith, S., L. Finzi, and C. Bustamante, 1992. Direct
mechanical measurements of the elasticity of single
DNA molecules by using magnetic beads. Science
258:1122 -- 1126.
[8] Schemmel, A., and H. E. Gaub, 1999. Single molecule
force spectrometer with magnetic force control and
inductive detection. Rev. Sci. Instrum. 70:1313 -- 1317.
[9] Gosse, C., and V. Croquette, 2002. Magnetic Tweez-
ers: Micromanipulation and Force Measurement at the
Molecular Level. Biophys. J. 82:3314 -- 3329.
[10] Neuman, K. C., and A. Nagy, 2008. Single-molecule
force spectroscopy: optical tweezers, magnetic tweezers
and atomic force microscopy. Nat. Methods 5:491 -- 505.
[11] Long, X., J. W. Parks, C. R. Bagshaw, and M. D. Stone,
2013. Mechanical unfolding of human telomere G-
quadruplex DNA probed by integrated fluorescence and
magnetic tweezers spectroscopy. Nucleic Acids Res.
41:2746 -- 2755.
[12] Rief, M., M. Gautel, F. Oesterhelt, J. M. Fernandez,
and H. E. Gaub, 1997. Reversible Unfolding of Individ-
ual Titin Immunoglobulin Domains by AFM. Science
276:1109 -- 1112.
[13] Oesterhelt, F., M. Rief, and H. E. Gaub, 1999. Single
molecule force spectroscopy by AFM indicates helical
structure of poly(ethylene-glycol) in water. New J. Phys.
1:6.1 -- 6.11.
[14] Rief, M., and H. Grubmüller, 2002. Force Spectroscopy
of Single Biomolecules. ChemPhysChem 3:255 -- 261.
9
[15] Puchner, E. M., and H. E. Gaub, 2009. Force and
function: probing proteins with AFM-based force spec-
troscopy. Curr. Opin. Struct. Biol. 19:605 -- 614.
[26] Harris, L. J., S. B. Larson, K. W. Hasel, and A. McPher-
son, 1997. Refined Structure of an Intact IgG2a Mono-
clonal Antibody † , ‡. Biochemistry 36:1581 -- 1597.
[16] Scholl, Z. N., Q. Li, and P. E. Marszalek, 2014. Sin-
gle molecule mechanical manipulation for studying
biological properties of proteins, DNA, and sugars. Wi-
ley Interdiscip. Rev. Nanomedicine Nanobiotechnology
6:211 -- 229.
[17] Nishizaka, T., H. Miyata, H. Yoshikawa, S. Ishiwata, and
K. Kinosita, 1995. Unbinding force of a single motor
molecule of muscle measured using optical tweezers.
Nature 377:251 -- 254.
[18] Kellermayer, M. S. Z., S. B. Smith, H. L. Granzier, and
C. Bustamante, 1997. Folding-Unfolding Transitions
in Single Titin Molecules Characterized with Laser
Tweezers. Science 276:1112 -- 1116.
[19] Tskhovrebova, L., J. Trinick, J. A. Sleep, and R. M.
Simmons, 1997. Elasticity and unfolding of single
molecules of the giant muscle protein titin. Nature
387:308 -- 312.
[20] Wang, M., H. Yin, R. Landick, J. Gelles, and S. Block,
1997. Stretching DNA with optical tweezers. Biophys.
J. 72:1335 -- 1346.
[21] Bennink, M. L., S. H. Leuba, G. H. Leno, J. Zlatanova,
B. G. de Grooth, and J. Greve, 2001. Unfolding individ-
ual nucleosomes by stretching single chromatin fibers
with optical tweezers. Nat. Struct. Biol. 8:606 -- 610.
[22] Wen, J.-D., M. Manosas, P. T. Li, S. B. Smith, C. Busta-
mante, F. Ritort, and I. Tinoco, 2007. Force Unfolding
Kinetics of RNA Using Optical Tweezers. I. Effects of
Experimental Variables on Measured Results. Biophys.
J. 92:2996 -- 3009.
[23] Oliver, R. C., D. J. Read, O. G. Harlen, and S. A. Harris,
2013. A stochastic finite element model for the dynamics
of globular macromolecules. J. Comput. Phys. 239:147 --
165.
[24] Richardson, R. A., K. Papachristos, D. J. Read, O. G.
Harlen, M. Harrison, E. Paci, S. P. Muench, and S. A.
Harris, 2014. Understanding the apparent stator-rotor
connections in the rotary ATPase family using coarse-
grained computer modeling. Proteins Struct. Funct.
Bioinforma. 82:3298 -- 3311.
[25] Hanson, B., R. Richardson, R. Oliver, D. J. Read,
O. Harlen, and S. Harris, 2015. Modelling biomacro-
molecular assemblies with continuum mechanics.
Biochem. Soc. Trans. 43:186 -- 192.
10
[27] Janeway, C. A., P. Travers, M. Walport, and M. Shlom-
chik, 2001. Immunobiology: The Immune System in
Health and Disease. Garland Science.
[28] Solernou, A., B. S. Hanson, R. A. Richardson, R. Welch,
D. J. Read, O. G. Harlen, and S. A. Harris, 2018. Fluc-
tuating Finite Element Analysis (FFEA): A continuum
mechanics software tool for mesoscale simulation of
biomolecules. PLOS Comput. Biol. 14:e1005897.
[29] Shontz, S. M., and S. A. Vavasis, 2012. A robust solution
procedure for hyperelastic solids with large boundary
deformation. Eng. Comput. 28:135 -- 147.
[30] Gent, A. N., 1996. A New Constitutive Relation for
Rubber. Rubber Chem. Technol. 69:59 -- 61.
[31] Fischer, H., I. Polikarpov, and A. F. Craievich, 2004.
Average protein density is a molecular-weight-dependent
function. Protein Sci. 13:2825 -- 2828.
[32] Quillin, M. L., and B. W. Matthews, 2000. Accurate
calculation of the density of proteins. Acta Crystallogr.
Sect. D Biol. Crystallogr. 56:791 -- 794.
[33] Oliver, R. C., 2013. A stochastic finite element model
for the dynamics of globular proteins. Ph.D. thesis,
University of Leeds.
[34] Voss, A., C. Dietz, A. Stocker, and R. W. Stark, 2015.
Quantitative measurement of the mechanical properties
of human antibodies with sub-10-nm resolution in a
liquid environment. Nano Res. 8:1987 -- 1996.
[35] Abraham, M. J., T. Murtola, R. Schulz, S. Páll, 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.
[36] Lindorff-Larsen, K., S. Piana, K. Palmo, P. Maragakis,
J. L. Klepeis, R. O. Dror, and D. E. Shaw, 2010. Improved
side-chain torsion potentials for the Amber ff99SB pro-
tein force field. Proteins Struct. Funct. Bioinforma.
78:1950 -- 1958.
[37] Abraham, M., B. Hess, D. van der Spoel, and E. Lindahl,
2012. GROMACS Reference Manual, version 5.1.2.
[38] Using FFEA to perform a full analysis of the torsional
rigidity of the molecule turns out to not be possible: for
high values of φ (resulting from high values of the torque
τ), the molecule relaxes its internal stresses by twisting
back the linker region and thereby moving through
its own surface, which is a non-physical phenomenon.
Unfortunately, incorporating a steric surface-surface
repulsion for the molecule in order to prevent this results
in computationally infeasible calculations at late times:
the linker region coils up and much of the surface
remains at a close distance to itself. The simulations
eventually fail because the molecule cannot relax its
internal stresses and as a result the elements in the linker
invert due to the strong forces on the nodes.
[39] Au, C. K., 2008. The Geometric Interpretation of Link-
ing Number, Writhe and Twist for a Ribbon. Comput.
Model. Eng. Sci. 29:151 -- 162.
[40] Humphrey, W., A. Dalke, and K. Schulten, 1996. VMD:
Visual molecular dynamics. J. Mol. Graph. 14:33 -- 38.
11
|
1203.4823 | 1 | 1203 | 2012-03-21T20:05:25 | Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks | [
"physics.bio-ph",
"cond-mat.soft"
] | Biopolymer networks are of fundamental importance to many biological processes in normal and tumorous tissues. In this paper, we employ the panoply of theoretical and simulation techniques developed for characterizing heterogeneous materials to quantify the microstructure and effective diffusive transport properties (diffusion coefficient $D_e$ and mean survival time $\tau$) of collagen type I networks at various collagen concentrations. In particular, we compute the pore-size probability density function $P(\delta)$ for the networks and present a variety of analytical estimates of the effective diffusion coefficient $D_e$ for finite-sized diffusing particles. The Hashin-Strikman upper bound on the effective diffusion coefficient $D_e$ and the pore-size lower bound on the mean survival time $\tau$ are used as benchmarks to test our analytical approximations and numerical results. Moreover, we generalize the efficient first-passage-time techniques for Brownian-motion simulations in suspensions of spheres to the case of fiber networks and compute the associated effective diffusion coefficient $D_e$ as well as the mean survival time $\tau$, which is related to nuclear magnetic resonance (NMR) relaxation times. Specifically, the Torquato approximation provides the most accurate estimates of $D_e$ for all collagen concentrations among all of the analytical approximations we consider. We formulate a universal curve for $\tau$ for the networks at different collagen concentrations. We apply rigorous cross-property relations to estimate the effective bulk modulus of collagen networks from a knowledge of the effective diffusion coefficient computed here. | physics.bio-ph | physics | Quantitative Characterization of the Microstructure
and Transport Properties of Biopolymer Networks
Yang Jiao1, Salvatore Torquato1,2,3,4,5
1 Physical Science in Oncology Center, Princeton Institute for the Science and
Technology of Materials, Princeton University, Princeton, NJ 08544, USA
2 Department of Chemistry, Princeton University, Princeton, NJ 08544, USA
3 Department of Physics, Princeton University, Princeton, NJ 08544, USA
4 Princeton Center for Theoretical Science, Princeton University, Princeton, NJ
08544, USA
5 Program in Applied and Computational Mathematics, Princeton University,
Princeton, NJ 08544, USA
Corresponding author contact information:
Salvatore Torquato
Tel.: 609-258-3341
Fax: 609-258-6746
E-mail: [email protected]
Short title: Microstructure and transport properties of collagen
Classification numbers: 87.15.rp, 87.85.jc
2
1
0
2
r
a
M
1
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
3
2
8
4
.
3
0
2
1
:
v
i
X
r
a
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks2
Abstract.
Biopolymer networks are of fundamental
importance to many biological
processes in normal and tumorous tissues.
In this paper, we employ the
panoply of theoretical and simulation techniques developed for characterizing
heterogeneous materials to quantify the microstructure and effective diffusive
transport properties (diffusion coefficient De and mean survival time τ ) of collagen
type I networks at various collagen concentrations.
In particular, we compute
the pore-size probability density function P (δ) for the networks and present a
variety of analytical estimates of the effective diffusion coefficient De for finite-
sized diffusing particles, including the low-density approximation, the Ogston
approximation, and the Torquato approximation. The Hashin-Strikman upper
bound on the effective diffusion coefficient De and the pore-size lower bound on the
mean survival time τ are used as benchmarks to test our analytical approximations
and numerical results. Moreover, we generalize the efficient first-passage-time
techniques for Brownian-motion simulations in suspensions of spheres to the case
of fiber networks and compute the associated effective diffusion coefficient De as
well as the mean survival time τ , which is related to nuclear magnetic resonance
(NMR) relaxation times. Our numerical results for De are in excellent agreement
with analytical results for simple network microstructures, such as periodic arrays
of parallel cylinders. Specifically, the Torquato approximation provides the most
accurate estimates of De for all collagen concentrations among all of the analytical
approximations we consider. We formulate a universal curve for τ for the networks
at different collagen concentrations, extending the work of Yeong and Torquato
[J. Chem. Phys. 106, 8814 (1997)]. We apply rigorous cross-property relations to
estimate the effective bulk modulus of collagen networks from a knowledge of the
effective diffusion coefficient computed here. The use of cross-property relations
to link other physical properties to the transport properties of collagen networks
is also discussed.
PACS numbers: 87.15.rp, 87.85.jc
Keywords: collagen network, microstructure, macromolecule diffusion, heterogeneous
media
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks3
1. Introduction
Biopolymer networks, such as the cross-linked bundles (or fibers) of collagen and fibrin
in the extracellular matrix (ECM), provide mechanical support for cells and serve as
the media for the transmission of many biomechanical/biochemical cues that regulate
cell motility, proliferation, differentiation and apoptosis [1, 2, 3, 4]. The diffusion and
absorption of various macromolecules in biopolymer networks are of crucial importance
to the regulation and metabolism of normal organs and to the delivery of drugs
in tumor tissues [5, 6]. Such biological processes are largely determined by the
composition and microstructure of the network, especially the complex pore space
between the fibers [7, 8]. Thus, knowledge of the effective transport and mechanical
properties of biopolymer networks is crucial in order to understand quantitatively the
aforementioned biological processes.
The microstructure of biopolymer networks and their associated transport
properties have been investigated by many researchers. For example, Ogston et al.
[9, 10] introduced the idea of "influence cylinders" associated with each fiber in the
system, which enables one to obtain the probability distribution of spherical pores
with different sizes δ, i.e., the pore-size probability density function P (δ) (also referred
to as the pore-size distribution function in literature; see the definition in Sec. 2). For
polymer networks composed of very long and stiff fibers, Ogston derived an analytical
expression of P (δ), which depends on the volume fraction and the characteristic
diameter of the fibers [9]. Other approaches used to ascertain pore-size statistics
include Fourier analyses of three-dimensional (3D) confocal microscopy images [11],
or statistical analysis of nearest points on collagen fibers obtained from confocal-
microscopic image stacks [12]. Recently, a method based on the direct analysis of
the entire 3D real-space network geometry from high-contrast confocal microscopy
data has been developed [13, 14]. Specifically, Lindstrom et al.
[14] represented the
collagen fibers in the networks as thinned skeletal center lines and the cross-links
are represented as nodes, which can be thought of as the "graph" representation
of a biopolymer network. These authors also employed an inverse reconstruction
method to characterize the microstructure of the collagen networks and investigated
the mechanical properties of the networks using finite-element analysis.
The determination of the effective diffusion coefficient De for polymer networks
dates back to the pioneering work of Johansson, Lofroth and coworkers [15, 16, 17,
18, 19]. Johansson et al. experimentally studied the diffusion of small monodisperse
polyethylene glycols [15] and nonionic micelles [16] in polymer systems and accurately
measured the long-time-limit self-diffusion coefficient (i.e., De) using a tracer technique
[17]. By considering the local diffusion of a particle around a single fiber, Johansson
et al.
[18] derived an analytical approximation of De which incorporates the
microstructural information of the pore space. Since their approach was based on
the key concept of the distribution of "influence cylinders" introduced by Ogston,
the approximation of De is henceforth referred to as the Ogston approximation. To
test the predictive capacity of their theory, Johansson and Lofroth [19] carried out
hard-sphere Brownian-motion simulations, in which the diffusing particles were hard
spheres and the fibers were considered to hinder the diffusion of the particles. Although
hydrodynamic effects were not taken into account in their simulations, the results were
shown to be in excellent agreement with experimental data and theoretical predictions
for a wide range of particle sizes [19]. Recently, the effects of the anisotropy of fiber
orientations [20, 21] and of the hydrodynamic interactions between the particle and
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks4
fibers [22] on the effective diffusion coefficient have also been investigated. We note
that by mathematical analogy, the problem of macromolecular diffusion in the pore
space exterior to the collagen fibers is equivalent to the electrical or thermal conduction
problem in the pore space with perfectly insulating fibers [23].
Very recently, it has been suggested that the powerful theoretical and simulation
techniques developed for characterizing the microstructure and effective properties
of random heterogeneous materials [23, 24, 25, 26] could be fruitfully employed to
model complex biological systems, especially malignant tumors and the associated host
microenvironment [27]. This idea has led to fruitful applications in the understanding
of the spatial organizations of abnormal cells in brain tumors [28].
In this paper, we further explore these techniques from the theory of
heterogeneous materials by investigating the microstructure and transport properties
of collagen type I networks (i.e., the most abundant collagen of the human body
found in tissue and bones, and therefore, called "type I"; see Fig. 1). There exist
analytical expressions that relate the effective transport and mechanical properties
of general heterogeneous materials to their microstructure via a variety of n-point
correlation functions; see Ref. [23] and references therein. This formalism has led to
the evaluation of effective transport and mechanical properties for a variety of classes of
microstructures, including dispersions of penetrable [29, 30, 31], impenetrable spheres
[32, 33, 34], oriented fibers [35, 36] and ellipsoid suspensions [37, 23], fluid-saturated
rock [38], and interpenetrating ceramic-metal composites [39].
In the case of collagen-like networks, we calculate for the first time a variety
of structural descriptors as well as the associated transport properties, such as the
effective diffusion coefficient De and mean survival time τ of a Brownian particle
assuming that the fiber interface is perfectly absorbing (i.e., the average time that a
Brownian particle spends in the solvent before it gets trapped by the fibers). The
mean survival time is related to the nuclear magnetic resonance (NMR) relaxation
times as discussed below. The latter transport property is intimately related to
the pore statistics [23, 40]. We also employ a variety of approximation schemes for
the effective diffusion coefficient De, including the low-density approximation [23],
Ogston approximation [18] and the Torquato approximation based on the perturbation
(phase-property contrast) expansion of De [41, 23]. These approximations incorporate
different levels of microstructural
information in terms of various lower-order
correlation functions that statistically describe the network microstructure. The
Hashin-Strikman upper bound on the effective diffusion coefficient De [42] and the
pore-size lower bound on the mean survival time τ [40, 43] are used as benchmarks to
test our analytical approximations as well as numerical simulations. Specifically, we
generalize the efficient first-passage-time techniques for Brownian-motion simulations
in suspensions of spheres [44, 45, 46, 47, 48] to the case of fiber networks and
compute the associated effective diffusion coefficient De and mean survival time τ . Our
numerical results of De are in excellent agreement for the analytical results of simple
network microstructures such as periodic arrays of parallel cylinders. Moreover, we
show that the Torquato approximation provides the most accurate estimate of De for
all concentrations among the employed approximation schemes. We also formulate a
universal curve for τ for different networks at different collagen concentrations, i.e., we
devise a way to scale τ in such a way that the scaled data for different collagen networks
collapse onto a single curve. Rigorous cross-property relations [23] are applied to
estimate the effective bulk modulus of collagen networks from a knowledge of the
computed effective diffusion coefficient.
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks5
The rest of the paper is organized as follows: In Sec. 2, we define the statistical
descriptors that will be used to characterize the network microstructures. In Sec. 3,
we provide analytical approximations and rigorous bounds for the effective properties
In addition, we discuss the first-passage-time simulation techniques for
De and τ .
network structures in detail.
In Sec. 4, we present the analytical and numerical
results of the correlation functions and the effective transport properties. In Sec. 5,
we estimate the effective bulk modulus of collagen networks using the cross-property
relations from the computed effective diffusion coefficients of the networks. In Sec.
6, we make concluding remarks, including the use of cross-property relations to link
other physical properties to the transport properties of collagen networks.
2. Network Microstructure Characterization
A collagen network can be considered to be a two-phase heterogeneous material
composed of a fiber phase and solvent phase (i.e., the pore space), which is exterior to
the fibers. A important feature of such a network microstructure is that both phases
percolate across the system, i.e., there is a continuous path between any two point of
the phase of interest that is entirely in the phase of interest, even when the volume
fraction of the fiber phase (fraction of space covered by the fibers) is very low. In this
section, we will introduce various statistical microstructural descriptors for a general
two-phase material, including the n-point correlation functions Sn and the pore-size
probability density function P (δ).
2.1. n-Point Correlation Functions
Consider a two-phase heterogeneous material in which each phase has volume fraction
φi (i = 1, 2), it is customary to introduce the indicator function I (i)(x) defined as
(1)
I (i)(x) =(cid:26) 1,
0,
x ∈ Vi,
x ∈ ¯Vi,
where Vi is the region occupied by phase i and ¯Vi is the region occupied by the other
phase. The statistical characterization of the spatial variations of the material involves
the calculation of the standard n-point correlation functions:
S(i)
n (x1, x2,··· , xn) =DI (i)(x1)I (i)(x2)···I (i)(xn)E ,
where the angular brackets h···i denote an ensemble average.
The quantity
S(i)
n (x1, x2, . . . , xn) also gives the probability of finding n points positioned at
x1, x2, . . . , xn all in phase i.
(2)
For statistically homogeneous materials, the n-point correlation function depends
not on the absolute positions but on their relative displacements, i.e.,
n (x1, x2,··· , xn) = S(i)
S(i)
(3)
for all n ≥ 1, where xij = xj − xi. Thus, there is no preferred origin in the system,
which in Eq. (3) we have chosen to be the point x1.
In particular, the one-point
correlation function is a constant everywhere, namely, it is equal to the volume fraction
φi of phase i, i.e.,
n (x12,··· , x1n),
S(i)
1 =DI (i)(x)E = φi,
(4)
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks6
which is the probability that a randomly chosen point in the material belongs to phase
i. For statistically isotropic materials, the n-point correlation function is invariant
n
under rigid-body rotation of the spatial coordinates. For n ≤ d, this implies that S(i)
depends only on the distances xij = xij (1 ≤ i < j ≤ n).
In general, an infinite set of Sn with n = 1, 2, 3... is required to completely
determine the microstructure and thus, the effective properties of a heterogeneous
material [23]. Specifically, the effective property of interest can written as an infinite
series involving integrals of such correlation functions.
In practice, a complete
knowledge of all of the Sn is never available. However, it has been shown that certain
approximations that can be regarded as resummations of the infinite series expansion
that incorporate lower-order Sn (e.g., S2, S3 and S4) can provide accurate estimates
of the effective properties [23]. We note that since only certain weighted integrals of
the correlation functions are needed, excellent approximations of effective properties
can be obtained even without computing all of the Sn explicitly. We will discuss these
approximations in detail in Sec. 3.1.
2.2. Pore-Size Probability Density Function
The pore-size probability density function P (δ) first rose to characterize the pore
space in porous media [43] and was then generalized to characterize any heterogeneous
material [23]. For a statistically homogeneous and isotropic material, P (δ)dδ gives the
probability that a randomly chosen point in the pore space lies at a distance between
δ and δ + dδ from the nearest point on the pore-solid interface. Since it is a probability
density function with dimensions of inverse length, we have P (δ) ≥ 0 for all δ, and it
normalizes to unity, i.e.,
Z ∞
0
P (δ)dδ = 1.
At extreme values of P (δ), we have that
P (0) = s/φ1, P (∞) = 0,
where s is the pore-solid interface area per unit volume and φ1 is the volume fraction
of the pore space. Therefore, s/φ1 is the interface area per unit pore volume. The
moments of P (δ), defined as
(5)
(6)
< δn >=Z ∞
0
δnP (δ)dδ,
(7)
provide useful characteristic length scales of the pore space in the material. Certain
lower-order moments of P (δ) also arise in bounds on the mean survival time τ [43, 40],
which we will discuss in Sec. 3.
It is very difficult to obtain analytical expression of P (δ) for a general polymer
network. For networks composed of very long and stiff polymer fibers, Ogston [9]
derived an expression for P (δ), i.e.,
P (δ) =
2φ2(δ + a)
a2
e−φ2(δ+a)2/a2
,
(8)
where φ2 = 1 − φ1 is the volume fraction of the fibers and a is the fiber radius. For
δ = 0, Eq. (8) gives
P (0) = s.
(9)
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks7
Comparing Eq. (9) and Eq. (6), it is clear that the Ogston expression for P (δ) can
only provide good estimates of it at very low fiber volume fractions, i.e., φ2 → 0
and φ1 = 1 − φ2 ≈ 1. In general, the Ogston expression will underestimate P (δ) at
intermediate δ, as we will show in Sec. 4.
Given any network microstructure, the associated P (δ) can be numerically
computed. Specifically, one generates many test points that are randomly distributed
in the pore space exterior to the fibers and compute the distances from each test point
to the nearest fiber surface. This amounts to finding the largest test sphere centered
at the test point that is entirely in the pore space. The resulting distances are binned
to obtain a probability density function, which is then normalized to give P (δ).
3. Mean Survival Time and Effective Diffusion Coefficient
3.1. Theoretical Techniques
3.1.1. Mean Survival Time Consider the steady-state problem of diffusion of
macromolecules which are absorbed upon contacting the network fibers. This implies
that the rate of production of the macromolecules per unit volume G is exactly
compensated by the rate of removal by the traps. Locally, the process is described by
the following Poisson equation [23]
D1∆c = −G in V1,
c = 0 on ∂V,
(10)
where c is concentration of the macromolecule, D1 is diffusion coefficient of the
macromolecule in the pore space V1 and ∂V is the pore-fiber interface. The boundary
condition that c = 0 on ∂V assumes a perfectly absorbing interface. i.e., a diffusion-
controlled reaction. This boundary condition can easily be relaxed to take into account
partially absorbing interfaces [23, 40].
An important quantity is the mean survival
time τ associated with a
macromolecule which is the average time that a diffusing macromolecule spends in
the pore space before it gets trapped by the fibers. In many medical applications,
the efficiency of a drug strongly depends on the ability of the drug macromolecules to
diffuse through the extracellular space mainly composed of collagen without getting
trapped by the fibers [5]. It is noteowrthy that nuclear magnetic resonance (NMR)
relaxation in porous media, a widely used technique for biomedical imaging, yields an
NMR relaxation times from which one can extract the mean survival time we consider
here [23].
The mean survival time τ is inversely proportional to the trapping constant γ,
i.e.,
τ =
1
γφ1D1
,
where γ is defined via
G = γD1 < c >
(11)
(12)
where < c > is the ensemble-averaged concentration field.
The optimal
lower bound on the mean survival time τ that incorporates
information on the pore space in terms of the first moment of the pore-size probability
density function is given by [40, 43]
< δ >2
D1
,
τ ≥
(13)
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks8
where < δ >= R ∞
0 δP (δ)dδ. It can be seen from Eq. (13) that τ D1 can provide an
estimate of the average pore size of the network. A corresponding upper bound on
the trapping constant can be obtain by substituting Eq. (11) into Eq. (13).
3.1.2. Effective Diffusion Coefficient Consider macromolecules diffusing between
the fibers that are not absorbed by the fibers. The Brownian motions of the
macromolecules are hindered by the fibers in the network which results in an effective
diffusion coefficient De smaller than that of the pure solvent in the pore space D1.
By mathematical analogy, the problem of macromolecular diffusion in the pore space
exterior to the fibers is equivalent to the electrical or thermal conduction problem
in the pore space with perfectly insulating fibers. Specifically, the local "flux" J(x)
of macromolecules is proportional to a local "intensity" E(x) which is the negative
gradient of the macromolecule concentration field c(x), i.e.,
where
J(x) = D(x) · E(x) = −D(x) · ∇c(x),
D(x) =(cid:26) D1I, x ∈ V1
otherwise
0,
(14)
(15)
and I is the unit second-order tensor. Under steady-state conditions with no source
and sink terms, the conservation of macromolecules requires that J(x) be solenoidal
[23], i.e.,
∇ · J(x) = 0.
(16)
If D(x) in Eq. (14) is replaced by the local conductivity tensor σ(x), one
obtains the local governing equations for conduction problems. We would like to
emphasize that although the diffusion problem and conduction problem are equivalent
in their mathematical formulations, there is an important distinction between the
effective diffusion coefficient De and the effective conductivity σe. For the conduction
problem, although the fiber phase is insulating, its contribution to σe is still explicitly
considered. For example, suppose that one randomly places test particles and tracks
their Brownian motions to compute σe. There is a fraction of total number of particles
φ2, which are initially in the insulating fiber phase and will be trapped there forever.
Clearly these test particles, which have a zero diffusion coefficient, are taken into
account in the ensemble average for σe. On the other hand, only the test particles in
the pore space are considered in order to compute De. Therefore, De and σe for the
same microstructure are related to one another via the following relation:
De
D1
=
1
φ1
σe
σ1
=
1
(1 − φ2)
σe
σ1
.
(17)
In the following, we present rigorous bounds and various analytical approxima-
tions for the effective diffusion coefficient De. These results were reported in literature
for the effective conductivity σe of a general heterogeneous material. Here, we modify
them according to Eq. (17) to obtain expressions for De.
Hashin-Strikman Upper Bound
For a two-phase heterogeneous material with an arbitrary but
isotropic
microstructure in which one of the phases (e.g., phase 2) is insulating, the Hashin-
Strikman (HS) upper bound for the effective diffusion coefficient De is given by
De
D1
=
2
2 + φ2
.
(18)
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks9
The HS lower bound in this case is trivially zero for all values of φ2 [23].
Although only volume fractions explicitly appear in the expression, it has been
shown that HS bound is the optimal bound given the two-point information of the
isotropic microstructure, i.e., S2 [23]. Specifically, it is shown that the HS bounds are
realizable for a special class of "coated sphere" model microstructures [23]. However,
it is clear that such two-point information is far from a complete characterization of
the network microstructure. Therefore, it can be expected that HS upper bound is
not tight, as we will show in Sec. 4.
Low-Density Approximation
For a heterogeneous material with microstructure composed of well-defined
inclusions (such as spheres, ellipsoids or cylinders) in a matrix, the effective properties
of the material can be written as a power series of the volume fraction φ2 of the
inclusions [23]. When φ2 is sufficiently small, i.e., in the low-density limit, truncating
the power series through the first-order in φ2 can provide a reasonable estimate of the
effective properties of interest.
For fiber networks, we consider that each fiber is an elongated prolate spheroid in
the "needle" limit. In such cases, the effective diffusion tensor for dilute suspensions
of orientated needles is given by
where
0
0
(De)22
0
0
(De)11
De =
(De)11 = (De)22 = D1[1 − φ2 + O(φ2
(D2)33
(De)33 = D1[1 + O(φ2
2)].
0
0
,
2)],
(19)
(20)
For statistically isotropic materials such as suspensions of randomly orientated needles
in a matrix, the effective diffusion coefficient is the average of the three principal
components of the tensor De, i.e.,
De
D1
= 1 −
2
3
φ2 + O(φ2
2).
(21)
Physically, Eq. (21) corresponds to the diffusion of Brownian particles in a
matrix with a single infinitely long fiber, which clearly underestimates De for actual
biopolymer networks. However, for certain microstructures such as periodic arrays
of parallel cylinders at low volume fractions, Eq. (21), which we call the low-density
approximation, can provide accurate estimates of De, which can be used as benchmarks
to test our simulation results.
Ogston Approximation
An improved approximation for De over the aforementioned low-density
approximation can be obtained if the contributions of multiple fibers are taken into
account simultaneously. This can be done by using the idea of an "influence cylinder"
associated with each fiber introduced by Ogston [10]. Specifically, consider a "coated
cylinder" with outer radius b and inner radius a (i.e., the radius of the fiber). One
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks10
can easily compute the local effective diffusion coefficient DL(b) associated with the
coated cylinder, i.e.,
DL(b)
D1
=
1
1 + a2/b2 =
1
1 + φ2(b)
,
(22)
where φ2(b) is local volume fraction of the fiber in the coated cylinder.
Now consider that the global De of a network is a weighted average of the local
DL(b) for the influence cylinders associated with each fiber, which leads to the relation
De
D1
=Z ∞
a
f (b)
DL(b)
D1
db,
(23)
where f (b) is the influence cylinder distribution function [18]. Ogston and coworkers
assume that the influence cylinders contribute to the global De in the same way they
contribute to the pore-size probability density function P (δ), i.e.,
P (δ) =Z ∞
a
f (b)g(b, δ)db,
(24)
where g(b, δ) is the local pore-size distribution associated with a coated cylinder with
outer radius b given by
g(b, δ) =
2(a + δ)
b2 H[δ − (b − a)]
(25)
and H(x) is the Heaviside step function, equal to unity for x > 0 and zero otherwise.
Then f (b) can be obtained by de-convolution of Eq. (24) with a knowledge of P (δ),
either from direct numerical sampling or theoretical considerations.
Torquato Approximation
Torquato has derived an expansion of the effective conductivity σe of any
two-phase heterogeneous materials in terms of the contrast (difference) between
the conductivities of the individual phases [41]. He showed that the [2,2] Pad´e
approximant of the so-called "strong-contrast" expansion, which incorporates up to
4-point microstructural information involving integrals over S2, S3 and S4 can provide
excellent estimates of σe for a wide range of model microstructures [41].
Here, we present the modified 4-point Pad´e approximation of the effective
diffusion coefficient De for collagen networks, which is henceforth referred to as the
Torquato approximation, i.e.,
De
D1
=
1
1 − φ2
(1 + 1
2
(1 + 1
2
γ2
ζ2 − 1
ζ2 − 1
γ2
2 ζ2) + (−1 + 1
2 ζ2) + ( 1
2 + 1
2 ζ2 − 1
2 ζ2 + 1
γ2
2
ζ2
γ2
ζ2
4
)φ2
)φ2
,
(26)
where the parameter ζ2 is a weighted integral that involves correlation functions S1,
S2 and S3 of the fiber phase; and the parameter γ2 is a weighted integral that involves
correlation functions S1, S2, S3 and S4 of the fiber phase. The readers are referred
to Ref. [41] for detailed discussions of these parameters. Useful rigorous inequalities
relating ζ2 and γ2 for three-dimensional microstructure are the following [41]:
− 1 ≤ γ2/ζ2 ≤ 1 − 2ζ2.
(27)
It is in general nontrivial to compute the 3-point and 4-point correlation functions
S3 and S4, even for relatively simple model microstructures (e.g., dispersions of
spheres), to obtain exact values of ζ2 and γ2. However, has been shown that
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks11
simplfied estimates of these parameters based on limited microstructural information
in conjunction with Eq. (26) can lead to excellent approximations for the effective
properties [41, 23]. Here, we consider the low-density approximation of ζ2 for a prolate
spheroid in the needle limit and only keep the leading order term, i.e., ζ2 = 1/4 [23].
The inequalities given in Eq. (27) then become
− 1 ≤ γ2/ζ2 ≤ 1/2.
(28)
It has been shown that for dispersions of spheres, γ2 = 0 can provide a very accurate
approximation formula for the effective conductivity of a wide range of sphere volume
fractions [41]. We will show in Sec. 4 that a proper choice of γ2 value that is near
its lower bound can provide an excellent approximation for the effective diffusion
coefficient associated with biopolymer networks.
3.2. First-Passage-Time Simulation Techniques
A straightforward way of numerically studying Brownian motion is to simulate the
exact zig-zag path of a diffusing particle (e.g., see Ref. [19]). However, it is clear that
this direct approach is not efficient means of obtaining effective diffusive properties,
since the details of the diffusion paths are averaged out and do not contribute to
the effective behavior. Moreover, one needs to consider a wide range of step sizes
associated with each random Brownian jump to extrapolate the results to the case of
infinitesimal small step size.
An alternative but much more computationally efficient approach is the first-
passage-time (FPT) simulation technique introduced by Torquato and coworkers
[44, 45, 46, 47, 48]. The key idea of the FPT approach is not to simulate the details
of the zig-zag diffusion paths but rather consider the average time that it takes a
Brownian particle to "jump" directly to a random location on the surface of the
largest imaginary sphere that is centered at the original position of the particle and
entirely within the solvent (i.e., the pore phase). The imaginary sphere is referred to
the "first-passage sphere" (FPS), whose radius is R. It can be shown that the mean
time τFPS for a Brownian particle, which is initially at the center of the FPS and takes
a complicated zig-zag path to hit the surface of the FPS is, in three dimensions, given
by
τFPS = R2/(2dD1),
where d = 3 is the space dimension.
(29)
When the particle is very close to the fiber surface, i.e., the distance r from the
particle centroid to the fiber surface is smaller than a prescribed tolerance ∆, we
consider that the particle hits the fiber surface and is reflected back. The FPS in this
case encloses both the pore phase and fiber phase. Suppose that the FPS centered at
the Brownian particle centroid possess a radius R, the associated time τREF that the
Brownian particle takes to hit the fiber surface, be reflected back and hit the FPS can
be estimated by
τREF(r) =
R2
6D1
V1 + V2
V1
"1 +
1
2(cid:16) r
R(cid:17)2
1
2
−
∞
Xm=0
C2m+1(cid:16) r
R(cid:17)2m+1# , (30)
where 0 ≤ r ≤ R, V1 and V2 are the volume of the pore phase and the volume of the
fiber phase enclosed in the FPS, respectively, and
C2m+1 =
(−1)m+1(2m)!
22m+1(m!)2
3(4m + 3)
(2m − 1)(m + 2)(m + 1)
.
(31)
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks12
Equation (30) was first derived by Torquato and coworkers [45, 46, 47], and it has been
shown to provide excellent approximation of the exact τREF for any local geometry
when r ≪ R and R is smaller than the diameter of the fiber. The readers are referred
to Ref. [45] and the references therein for additional details.
To compute τREF using Eq. (30), a key step is to evaluate the intersection volume
between a sphere and a cylinder (i.e., V2), the details of which are given in the
Appendix. In the rare case that the particle is close to a cross link (junction of several
fibers), V2 is the volume of the fiber junction enclosed in the FPS, which is computed
by Monte Carlo sampling [47]. For example, one randomly places test points in the
FPS and computes the fraction of times that the point falls into the vicinity of the
fiber junction.
To obtain De, one considers an ensemble of Brownian trajectories in the pore
space. When a diffusing particle is sufficiently far away from the fiber surface, one
constructs the largest FPS of radius R around the diffusing particle which just touches
the fiber surface. The particle then jumps in one step to a random point on the surface
of the FPS and the process is repeated, each time keeping track of R2
i , until the particle
is within a prescribed very small distance ∆ to the fiber surface (see Fig. 3). At this
point in time, the particle is considered to hit the fiber and then is reflected back.
Thus, one keeps track of the radius Rj of the FPS that encloses both the fiber phase
and the pore phase and computes the associated time τREF(Rj). The expression for
effective diffusion coefficient De is then given by
De
D1
=*
Pi Ri + 6D1Pj τREF(Rj )+ ,
Pi Ri +Pj Rj
(32)
where τREF(R) is given by Eq. (30) and < . > denotes the ensemble average over many
Brownian particles. In our simulations, we use N = 5 000 Brownian particles.
The mean survival time τ can be obtained in a similar way [44]. Specifically, one
constructs the first-passage-time path composed of many jumps to the surface of FPS
associated with a Brownian particle and keeps track of Ri for each FPS. When the
particle is within ∆ to the fiber surface, it is considered trapped by the fiber. Thus,
the mean survival time can be computed via
τ =*Xi
Ri/D1+ ,
where < . > denotes ensemble average over many Brownian particle.
simulations, we use N = 5 000 Brownian particles.
(33)
In our
We note that in the aforementioned first-passage-time simulation technique, we
consider that the Brownian particle is "point" particle with zero diameter. For finite-
sized particles diameter dP, it has been shown that one can still consider "point"
particles in a network microstructure with the diameter of the fibers dF dilated by dP
[48, 49].
4. Results
Our data are "graph representations" of collagen type I networks [14] with final
collagen concentrations of 1.0 mg/ml, 2.0 mg/ml and 4.0 mg/ml. The fibers roughly
possess a circular cross section of diameter dF = 1.0 × 10−7 m [14]. The average fiber
lengths ℓF for the networks with the three collagen concentrations are respectively
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks13
1.96 × 10−6m, 1.81 × 10−6m, and 1.28 × 10−6m. The corresponding volume fractions
of the fibers are respectively 1.7 × 10−3, 2.4 × 10−3, and 5.2 × 10−3. The tolerance
∆ described in Sec. 3 is chosen to be ∆ = 5 × 10−3dF = 5.0 × 10−10 m. Since we
do not consider hydrodynamic effects in our simulations, we only consider particles
with diameter dP comparable to the fiber diameter dF, i.e., dP ≤ dF. For large dP,
it has been shown that the hydrodynamic effects on Brownian motion are significant
[22]. The results reported below are ensemble averages of three independent network
configurations at each collagen concentration.
4.1. Pore-Size Probability Density Function
The pore-size probability density functions P (δ) for the collagen network at three
concentrations are numerically computed as described in Sec. 2. The obtained P (δ)
are shown in Fig. 4 and compared to the corresponding Ogston expressions [Eq. (8)]
at the same fiber volume fractions.
It can be clearly seen that the Ogston expression of P (δ) overestimates the number
of intermediate pores and underestimates the number of large pores in the system.
This is because in the derivation of Eq. (8), it is assumed that the network is composed
of fibers with very long persistence length. For the collagen networks we study, the
average fiber lengths ℓ are less than twice of the corresponding averaged pore size
< δ > [defined in Eq. (7)], which are respectively 1.22 × 10−6m, 0.998 × 10−6m,
and 0.684× 10−6m for collagen concentrations 1.0 mg/ml, 2.0 mg/ml and 4.0 mg/ml.
Therefore, the long-fiber-length assumption for the Ogston expression is not true here.
The P (δ) data will be employed to compute the lower bound on the mean survival
time τ [Eq. (13)] and to compute the Ogston approximation for the effective diffusion
coefficient De [Eq. (23)] in the following sections.
4.2. Mean Survival Time
The mean survival time τ is computed using the first-passage-time technique described
in Sec. 2. Figure 5 shows the scaled dimensionless mean survival time τ D1/ℓ2
F for
the collagen networks with three different concentrations as a function of Brownian
particle diameter. It can be seen that as the collagen concentration increases, larger
particles are more easily get trapped by the fibers. This fact is of great importance in
cancer chemotherapy, which we will discuss in Sec. 6.
As indicated in Sec. 3, the diffusion of finite-sized particles in the original network
is equivalent to the diffusion of point particles in a properly dilated network, which
possesses a higher fiber volume fraction. Figure 6 shows the scaled mean survival time
τ D1/ℓ2
F for the collagen networks with three different concentrations as a function
of the particle diameter. The pore-size lower bounds are also shown.
It can be
seen that although the bounds are not sharp, they do not deviate very much from
the actual mean survival times. We note that these bounds only incorporate partial
information about the pore-size probability density function P (δ), namely, the first
moment < δ > of P (δ). Therefore, one would expect that incorporating the full
information content of P (δ) would lead to good predictions of the effective diffusive
properties considered here. Indeed, we will show in the following section that the a
generalization of the Ogston approximation that employs the complete microstructural
information contained in P (δ) provides a very good estimate of De.
In Ref. [50], Torquato and Yeong found a universal curve for a scaled mean
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks14
survival time τ for a wide range of microstructures with different porosities, including
various random and ordered distributions of spheres and certain continuous models.
Specifically, the universal curve has the following form:
τ
τ0
= a1x + a2x2,
where a1 and a2 are constants and
τ0 =
3φ2
D1φ1s2 ,
x =
< δ >2
τ0D1
,
(34)
(35)
and s is the specific surface, i.e., the solid-pore interface area per unit volume. For
the class of microstructures they studied, Torquato and Yeong found that a1 = 8/5
and a2 = 8/7.
For the collagen networks studied here, we find that Eq. (34) also holds (see
Fig. 7). However, the constants are different from those obtained by Torquato and
Yeong, i.e, we find that a1 = 0.121 and a2 = 1.88 for the networks with different
concentrations. A possible reason for the difference in the constants is that collagen
networks do not belong to the same class of microstructures studied in Ref. [50], which,
for example, do not contain filamentary-like structures, as in the case of collagen fibers.
This implies that there could exist a more general scaling curve for the mean survival
time that incorporates both the networks and the microstructures studied in Ref. [50].
Nonetheless, our results enable one to efficiently estimate the properties of collagen
networks. In particular, given any of the three quantities among the four quantities
τ , φ1, s and < δ >, the remaining one can be estimated employing Eq. (34).
4.3. Effective Diffusion Coefficient
The effective diffusion coefficient De for various network microstructures are computed
using both the theoretical techniques described in Sec. 3.1 and the first-passage-time
technique described in Sec. 3.2. Figure 8 shows De for the fiber networks with different
collagen concentrations as a function of the Brownian particle diameter. Similar to the
case of the mean survival time, as the collagen concentration φ increases, it becomes
more and more difficult for larger particles to diffuse in the collagen.
Figure 9 shows De as a function of the fiber volume fraction.
In addition to
the results for the collagen networks, we also show De for a model microstructure
composed of parallel cylinders arranged on a square lattice. The Hashin-Strikman
(HS) upper bound and various approximations of De discussed in Sec. 3.1 are also
shown in Figure 9. As we indicated earlier, since the HS bound only incorporates
the limited two-point information S2,
it can not provide a good estimate of De.
This is also evident from the fact that the HS bound is realized by certain class of
"coated sphere" model microstructures, which are clearly topologically distinct from
the network microstructures because one of the phases is topologically disconnected.
It can be seen that the low-density approximation also underestimates De for the
collagen networks. This is because it only considers the effect of a single long fiber
to the diffusing particles. In the actual networks, the average fiber length is less than
twice the average pore size as we indicated earlier. Moreover, the cross-links also
significantly hinder the diffusion of the particles. However, for the parallel-cylinder
model at low volume fractions, the low-density approximation should provide accurate
estimates, since in such cases, the diffusion of the particles is only hindered by well
separated single cylinders. Indeed, we find that the approximation agrees very well
with our simulation data, which also verifies the accuracy of our simulations.
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks15
The Ogston approximation that incorporates the pore-size information of the
networks provides a much better estimate of De compared to the HS bound and the
dilute approximation. However, it still slightly underestimates De for large particles
in networks at high collagen concentrations. This is because the "influence cylinders"
(see Sec. 3.1) are associated with individual long fibers and the effects of finite-fiber
length and the cross-links are still not fully incorporated. On the other hand, one can
see that the Torquato approximation agrees extremely well with the simulation data
for all volume fractions that we considered. Specifically, in employing Eq. (26) we have
chosen the 4-point parameter value such that γ2/ζ2 = −0.925. Note that this value is
very close to the lower bound value -1, which is not very surprising the networks can
be considered as a kind of "limit" microstructure. This does not mean the actual value
of γ2/ζ2 if computed with a full knowledge of the associated 3-point function S3 and 4-
point function S4, since the value of ζ2 we used is also an approximation. The success
of the Torquato approximation is due to the fact that higher-order microstructural
information that possibly reflects the effects of the cross-links is already taken into
account by the 4-point parameter γ2 in the expression Eq. (26).
5. Estimating Elastic Properties of Collagen Network Using
Cross-Property Relations
Since effective properties of heterogeneous materials reflect certain microstructural
information about the material, it is possible to extract rigorously information about
one physical property given an accurate determination of a different effective property
obtained either experimentally or theoretically. Such interrelationships are called
cross-property relations [23, 56, 57, 58, 59, 60]. Rigorous cross-property relations
become especially useful if one property is more easily measured than another property.
In this section, we estimate the effective bulk modulus Ke [23] of fluid-saturated
collagen networks using the effective diffusion coefficient De computed here using the
cross-property relations. In particular, Gibiansky and Torquato [56, 57, 58] derived a
nontrivial rigorous cross-property upper bound K U
e on the effective bulk modulus Ke
of a fluid-saturated porous material with an insulating solid phase given the effective
conductivity (equivalent to the effective diffusion coefficient) of the material, i.e.,
Ke ≤ K U
e = K1e −
2G2(K1 − K2)2
2φ1φ2
a[3aφ1F − 3a − 2φ2G2]
,
where
a = φ2K1 + φ1K2 + 4G2/3,
K1e = φ1K1 + φ2K2 − φ1φ2(K1 − K2)2/a,
(36)
(37)
and φ1 and K1 are respectively the volume fraction and bulk modulus of the fluid
phase; φ2, K2, G2 are respectively the volume fraction, bulk modulus and shear
modulus of the solid phase. Moreover, the "formation factor" F in Eq. (36) is given
by
F =
1
φ1
D1
De
(38)
where D1 is the diffusion coefficient of the fluid phase and De is the effective diffusion
coefficient of the porous material.
For the collagen networks considered here, the solid phase corresponds to the
collagen fibers. The bulk modulus Ke of fluid-saturated collagen networks at fiber
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks16
volume fraction φ2 = 0.005 has been measured experimentally [61], i.e., Ke ≈ 2500
Pa. The shear modulus of "dry" collagen networks (i.e.., fiber networks without fluid)
have been computed numerically by Lindstrom et al.
[14], i.e., G2 = 24 Pa. We
use K2 = 10 Pa for the "dry" network and K1 = 2 GPa for the fluid. Using the
Torquato approximation for the effective diffusion coefficient De, the upper bound
value obtained from inequality (36) is computed, i.e., K U
e = 3530 Pa, which provides
a surprisingly good estimate of Ke, given the fact that K U
e is a rigorous upper bound
[23].
6. Conclusions and Discussion
In this paper, we have quantitatively characterized the microstructure, the mean
survival time τ , and the effective diffusion coefficient De of collagen type I networks by
applying theoretical and computational techniques from the theory of heterogeneous
materials. Specifically, we have computed the pore-size probability density function
P (δ) for the networks. We have also employed a variety of theoretical approximation
schemes for the effective conductivity of a two-phase material to estimate the effective
diffusion coefficient De for the networks. Such estimates include the low-density
approximation, the Ogston approximation, and the Torquato approximation, all of
which incorporate different levels of microstructural information about the networks.
The Hashin-Strikman upper bound on De and the pore-size lower bound on τ are used
as benchmarks to test our results. Moreover, we have generalize the efficient first-
passage-time techniques for Brownian-motion simulations in suspensions of spheres to
the case of network microstructures and compute the associated De and τ . We have
found a universal curve for τ for the networks at different collagen concentrations
and have shown that the Torquato approximation which takes into account higher-
order microstructural information can provide the most accurate estimate of De
for all collagen concentrations among the employed approximation schemes. Our
work also demonstrates that employing the rich family of theoretical and simulation
techniques developed in materials sciences to characterize biological systems (e.g.,
the heterogeneous host microenvironment of tumors) suggested in Ref. [27] is a very
promising approach worthy further exploration.
We have found that as the collagen concentration increases, the diffusion of large
particles in the collagen networks, and thus the extracellular matrix (ECM) becomes
more and more difficult and it is easier for the diffusing particles to be trapped by the
fibers. This is a major problem associated with any cancer chemotherapy, since drug
macromolecules would get trapped by collagen fibers without successfully diffusing to
the target site. It is known that a growing malignant tumor constantly modifies the
chemical composition of the collagen networks composing its ECM [51]. In addition,
since a pressure is built up as the tumor grows, the surrounding ECM is pushed
and compressed, leading to a higher collagen concentration in tumor ECM than in
normal tissues [52, 53, 54]. Therefore, it can be expected that the diffusion of drugs to
the tumors is really difficult. More efficient chemotherapies trying to overcome these
difficulties are being developed [55].
We also applied a rigorous cross-property upper bound to to estimate the effective
bulk modulus Ke of collagen networks from a knowledge of the effective diffusion
coefficient De computed here. The estimated value of Ke agrees well with existing
experimental data, given the fact that it is a rigorous upper bound.
In future work, we intend to generalize our simulation techniques and theoretical
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks17
approaches to investigate transport properties of tissues with both collagen networks
and various types of cells. Specifically, we will focus on the effects of the cell shape and
the plasma membrane on the diffusion of macromolecules. In addition, we will model
the mechanical behavior of tissues using the well-developed methods for heterogeneous
materials. Progress in these studies should deepen our understanding of the effects of
the host microenvironment on tumor growth and would lead to better cancer treatment
strategies.
In addition, we will apply cross-property relations to estimate other physical
properties of collagen networks from a knowledge of the effective diffusive transport
properties computed in this paper. In particular, given the latter, we will bound the
fluid permeability [59, 60] for the collagen networks studied here.
Acknowledgments
The authors are very grateful to S. B. Lindstrom and D. Weitz for providing the
data of the collagen networks. This project was supported by the National Center for
Research Resources and the National Cancer Institute of the National Institutes of
Health through Grant Number U54CA143803. The content is solely the responsibility
of the authors and does not necessarily represent the official views of the National
Cancer Institute or the National Institutes of Health.
Appendix: Intersection volume of a sphere with a cylinder
Consider a sphere with radius Rs centered at the surface of a cylinder with radius Rc,
the intersection volume VI of the sphere and the cylinder for the case Rc > Rs is given
by [62]
πR3
s+
VI =
[K(k)(A − B)(3B − 2A) + E(k)A(2A − 4B)] , (39)
where K(k) and E(k) are elliptic integrals of the first and second kind, respectively,
i.e.,
2
3
4
9√A
and
K(k) =Z 1
0
dz
p(1 − z2)(1 − k2z2)
, E(k) =Z 1
0
dzr 1 − k2z2
1 − z2 ,
A = 4R2
c, B = R2
S,
k2 = B/A.
(40)
(41)
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks18
Abbreviations list
• ECM: extracelluar matrix
• HS: Hashin-Strikman
• NMR: nuclear magnetic resonance
• FPS: first-passage sphere
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks19
References
[1] Kuntz RM and Saltzman WM 1997 Neutrophil motility in extracellular matrix gels: mesh size
and adhesion affect speed of migration. Biophys. J. 72 1472-80
[2] Lo CM, Wang HB, Dembo M and Wang YL 2000 Cell movement is guided by the rigidity of the
substrate. Biophys. J. 79 144-52
[3] Bischofs IB and Schwarz US 2003 Cell organization in soft media due to active mechanosensing.
Proc. Natl. Acad. Sci. 100 9274-79
[4] Grinnell F 2003 Fibroblast biology in three-dimensional collagen matrices. Trends Cell Biol. 13
264-69
[5] Comper WD 1996 Extracellular Matrix. (Amsterdam: Harwood Academic Publishers)
[6] Gevertz JL and Torquato S 2008 A novel three-phase model of brain tissue microstructure. PLoS
Comput. Biol. 4 e1000152.
[7] Jain RK 1987 Transport of molecules in the tumor interstitium: a review. Cancer Res. 47
3039-51 (1987).
[8] Yang YL, Motte S and Kaufman LJ 2010 Pore size variable collagen gels and their interaction
with glioma cells. Biomaterials 31 5678-88.
[9] Ogston AG 1958 The spaces in a uniform random suspension of fibres. Trans. Faraday Soc. 54,
1754-57
[10] Ogston AG, Preston BN and Wells JD 1973 On the transport of compact particles through
solutions of chain-polymers. Proc. R. Soc. London A 333 297-316
[11] Takahashi AR, et al. 2003 Real space observation of three-dimensional network structure of
hydrated fibrin gel. Colloid Polym. Sci. 281 832-38
[12] Kaufman LC, et al. 2005 Glioma expansion in collagen I matrices:
analyzing collagen
concentration-dependent growth and motility patterns. Biophys. J. 89 635-50 (2005).
[13] Mickel W, et al. 2008 Robust pore size analysis of filamentous networks from three-dimensional
confocal microscopy. Biophys. J. 95 6072-80.
[14] Lindstrom SB, Vader DA, Kulachenko A and Weitz DA 2010 Biopolymer network geometries:
characterization, regeneration, and elastic properties. Phys. Rev. E 82 051905
[15] Johansson L, Skanzte U and Lofroth JE 1991 Diffusion and interaction in gels and solutions. 2.
Experimental results on the obstruction effect. Macromolecules 24 6019-23
[16] Johansson L, Hedberg P and Lofroth JE 1993 Diffusion and interaction in gels and solutions. 4.
Hard sphere Brownian dynamics simulations. J. Phys. Chem. 97 747-55
[17] Johansson L and Lofroth JE 1991 Diffusion and interaction in gels and solutions I. Method.
Colloid Interface Sci. 142 116-20
[18] Johansson L, Elvingson C and Lofroth JE 1991 Diffusion and interaction in gels and solutions.
3 Theoretical results on the obstruction effect. Macromolecules 24 6024-9
[19] Johansson L and Lofroth JE 1993 Diffusion and interaction in gels and solutions. 5. Nonionic
micellar systems. J. Phys. Chem 98 7471-9
[20] Leddy HA, Haider MA and Guilak F 2006 Diffusional anisotropy in collagenous tissues:
fluorescence imaging of continuous point photobleaching. Biophys. J. 91 311-16
[21] Erikson A, et al. 2008 Physical and chemical modifications of collagen gels: impact on diffusion.
Biopolymers 89 135-43
[22] Stylianopoulos T, Diop-Frimpong B, Munn LL and Jain RK 2010 Diffusion Anisotropy in
Collagen Gels and Tumors: The Effect of Fiber Network Orientation Biophys. J. 99 3119-28
[23] Torquato S 2002 Random Heterogeneous Materials: Microstructure and Macroscopic Properties
(New York: Springer-Verlag)
[24] Sahimi M 2003 Heterogeneous Materials. Volume I: Linear Transport and Optical Properties
(New York: Springer-Verlag)
[25] Sahimi M 2003 Heterogeneous Materials. Volume II: Nonlinear and Breakdown Properties, and
Atomistic Modelling (New York: Springer-Verlag)
[26] Zohdi TI and Wriggers P (2005) Introduction to Computational Micromechanics. (New York:
Springer-Verlag)
[27] Torquato S 2011 Toward an Ising model of cancer and beyond. Phys. Biol. 8 015017
[28] Jiao Y, Berman H, Kiehl T and Torquato S 2011 Spatial organization and correlations of cell
nuclei in brain tumors. PLoS One 6 e27323
[29] Torquato S 1984 Bulk properties of two-phase media. I. Cluster expansion for the dielectric
constant of dispersions of fully penetrable spheres. J. Chem. Phys. 81 5079-88
[30] Torquato S 1985 Bulk properties of two-phase disordered media. II. Effective conductivity of a
dilute dispersion of penetrable spheres. J. Chem. Phys. 83 4776-85
[31] Torquato S 1986 Bulk properties of two-phase disordered media. III. New bounds on the effective
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks20
conductivity of dispersions of penetrable spheres. J. Chem. Phys. 84 6345-59
[32] Lado F and Torquato S 1986 Effective properties of two-phase disordered composite media: I.
Simplification of bounds on the conductivity and bulk modulus of dispersions of impenetrable
spheres. Phys. Rev. B 33 3370-78
[33] Lado F and Torquato S 1986 Effective properties of two-phase disordered composite media: II.
Evaluation of bounds on the conductivity and bulk modulus of dispersions of impenetrable
spheres. Phys. Rev. B 33 6428-34
[34] Beasley JD and Torquato S 1986 Bounds on the conductivity of a suspension of random
impenetrable spheres. J. Appl. Phys. 60 3576-81
[35] Torquato S and Lado F 1988 Bounds on the effective transport and elastic properties of
cylindrical fibers in a matrix. J. Appl. Mech. 55 347-54
[36] Miller CA and Torquato S 1991 Diffusion-controlled reactions among spherical traps: effect of
polydispersivity in trap size. J. Appl. Phys. 69 1948-55
[37] Lado F and Torquato S 1990 Two-point probability function for distributions of oriented hard
ellipsoids. J. Chem. Phys. 93 5912-17
[38] Coker D, Torquato S and Dunsmuir J 1996 Morphological and physical properties of
Fountainebleu sandstone from tomographic analysis. J. Geophys. Res. 101 17497-17506
[39] Torquato S, Yeong C, Rintoul MD, Milius D and Aksay IA 1999 Characterizing the structure
and mechanical properties of interpenetrating multiphase cermets. J. Amer. Cera. Soc. 82
1263-68
[40] Torquato S and Avellaneda M 1991 Diffusion and reaction in heterogeneous media: Pore size
distribution, relaxation times, and mean survival time. J. Chem. Phys. 95 6477-89
[41] Torquato S 1985 Effective electrical conductivity of two-phase disordered composite media. J.
Appl. Phys. 58 3790-97
[42] Hashin Z and Strikman S 1962 A variational approach to the theory of the effective magnetic
permeability of multiphase aaterials. J. Appl. Phys. 33 3125-32
[43] Prager S 1963 Interphase transfer in stationary two-phase media. Chem. Eng. Sci. 18 227-31
[44] Torquato S and Kim IC 1989 Efficient simulation technique to compute effective properties of
hetergeneous media. Appl. Phys. Lett. 55 1847-49
[45] Kim IC and Torquato S 1990 Determination of the effective conductivity of heterogeneous media
by Brownian motion simulation. it J. Appl. Phys. 68 3892-903
[46] Kim IC and Torquato S 1991 Effective conductivity of suspensions of hard spheres by Brownian
motion simulation J. Appl. Phys. 69 2280-89
[47] Kim IC and Torquato S 1992 Diffusion of finite-sized Brownian particles in porous media. J.
Appl. Phys. 71 2727-35
[48] Kim IC and Torquato S 1992 Diffusion of finite-sized Brownian particles in porous media. J.
Chem. Phys. 96, 1498-1503
[49] Torquato S 1991 Trapping of finite-sized Brownian particles in porous media. J. Chem. Phys.
95 2838-41
[50] Torquato S and Yeong CLY 1997 Universal scaling for diffusion-controlled reactions among traps.
J. Chem. Phys. 106 8814-20
[51] Hanahan D and Weinberg RA 2000 The hallmarks of cancer. Cell 100 57-70
[52] Helmlinger G, Netti PA, Lichtenbeld HC, Melder RJ and Jain RK 1997 Solid stress inhibits the
growth of multicellular tumor spheroids. Nature Biotech. 15 778-83
[53] Jiao Y and Torquato S 2012 Diversity of dynamics and morphologies of invasive solid tumors.
AIP Advances (in press)
[54] Jiao Y and Torquato S 2011 Emergent behaviors from a cellular automaton model for invasive
tumor growth in heterogeneous microenvironments. PLos Comput. Biol. 7 e1002314
[55] Joensuu H 2008 Systemic chemotherapy for cancer: from weapon to treatment. Lancet Oncol.
9 304
[56] Gibiansky LV and Torquato S 1993 Link between the conductivity and elastic moduli of
composite materials. Phys. Rev. Lett. 71 2927-30
[57] Gibiansky LV and Torquato S 1996 Link between the conductivity and elastic moduli of
composite materials. Proceedings of the Royal Society London A 452 253-83
[58] Gibiansky LV and Torquato S 1998 Rigorous connection between physical properties of porous
rocks. J. Geophys. Res. 103 23911-23
[59] Torquato S 1990 Relationship between permeability and diffusion-controlled trapping constant
of porous media. Phys. Rev. Lett. 64, 2644-46
[60] Avellaneda M and Torquato S 1991 Rigorous link between fluid permeability, electrical
conductivity, and relaxation times for transport in porous media. Phys. Fluids A 3, 2529-
40 (1991).
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks21
[61] Raub CB, Putnama AJ, Tromberg BJ and George SC 2010 Predicting bulk mechanical properties
of cellularized collagen gels using multiphoton microscopy. Acta Biomaterialia 6, 4657-4665
[62] Lamarche F and Leroy C 1990 Evaluation of the volume of intersection of a sphere with a
cylinder by elliptic integrals. Comput. Phys. Comm. 59 359-69
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks22
Figure Legends
• Figure 1: Collagen networks.
(a) Confocal microscope image of collagen-I
at a final collagen concentration 2.0mg/ml. The linear size of the image is
approximately 150 µm. Image courtesy of S. B. Lindstrom. (b) Three-dimensional
"graph" representation of the collagen network studied here. The linear size of
the box is approximately 100 µm.
• Figure 2: An illustration of sampling P (δ) of the collagen network in two
dimensions. Shown are the thin fibers that can possibly intersect as well as a
test point (red) and its associated largest sphere (blue) entirely within in the
pore space exterior to the fibers.
• Figure 3: An illustration of the first-passage-time simulation technique in two
dimensions. Shown are the thin fibers that can possibly intersect and several first
passage spheres. Starting from an initial position, a diffusing particle jumps to a
random location on the surface of its associated first-passage sphere. The process
is repeated until the particle is very close to the fiber surface. The the particle
will hits the fiber and be reflected back to the pore space.
at different collagen concentrations.
• Figure 4: The pore-size probability density function P (δ) for collagen networks
• Figure 5: The scaled dimensionless mean survival time τ D1/ℓ2
F for the collagen
networks at different collagen concentrations as a function of the diffusing particle
diameter.
• Figure 6: The scaled dimensionless mean survival time τ D1/ℓ2
F for the collagen
networks at different collagen concentrations as a function of the diffusing particle
diameter and the associated pore-size lower bound.
• Figure 7: Universal curve for the scaled mean survival time τ /τ0 versus < δ >2
/(τ0D1) for the collagen networks at different collagen concentrations. It is seen
that the scaled mean survival time for different collagen networks collapse onto a
single curve.
• Figure 8: The dimensionless effective diffusion coefficient De/D1 for the collagen
networks at different collagen concentrations as a function of the diffusing particle
diameter.
• Figure 9: The dimensionless effective diffusion coefficient De/D1 for the
collagen networks at different collagen concentrations as a function of the fiber
volume fraction. The Hashin-Strikman upper bound and various analytical
approximations as discussed in Sec. 3.1 are shown and compared to the simulation
data.
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks23
Glossary
• Collagen:
a group of naturally occurring and the most abundant proteins
(biopolymers) found in animals, especially in the flesh and connective tissues
of mammals.
• Heterogeneous material :
a material composed different materials (e.g., a
composite) or the same material in different state (e.g., a polycrystal). The fluid-
saturated collagen networks studied here are special heterogeneous materials.
• Diffusion coefficient : a proportionality constant between the molar flux due to
molecular diffusion and the gradient in the concentration of the species or the
driving force for diffusion.
• Mean survival time: the average time that a diffusing molecule spends in the
fluid phase before it gets trapped at the interface of the collagen fibers assuming
a perfectly absorbing interface.
• Bulk modulus: a measure of a material's resistance to uniform compression,
defined as the ratio of the infinitesimal pressure increase to the resulting relative
decrease of the material's volume.
• Shear modulus: a measure of a material's resistance to shape deformation shear
stress, defined as the ratio of shear stress to the shear strain.
a statistical descriptor of the microstructure of a
• Correlation function:
heterogeneous material, quantifying the spatial correlations at different points
in the microstructure.
• First-passage-time:
the average time that it takes for a diffusing particle to
"jump" directly to a random location on the surface of an imaginary sphere that
is centered at the original position of the particle and entirely within the solvent
region.
• Cross-property relation: an interrelationship that enables one to extract rigorously
information about one physical property of a heterogeneous material given an
accurate determination of a different property.
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks24
(a)
(b)
Figure 1. Jiao and Torquato
Figure 2. Jiao and Torquato
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks25
Figure 3. Jiao and Torquato
1mg/ml data
Ogston approx
1.5
1
0.5
)
δ
(
P
1.5
1
0.5
)
δ
(
P
2mg/ml data
Ogston approx
1.5
1
0.5
)
δ
(
P
4mg/ml data
Ogston approx
0
0
1
2
δ (µm)
(a)
3
4
0
0
1
2
δ (µm)
(b)
3
4
0
0
1
3
4
2
δ (µm)
(c)
Figure 4. Jiao and Torquato
1
0.75
F
1mg/ml
2mg/ml
4mg/ml
l
/
1
2
D
τ
0.5
0.25
0
0.5
dP/dF
1
1.5
Figure 5. Jiao and Torquato
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks26
1.5
1
F
1mg/ml data
pore-size bound
D
τ
2
l
/
1
1.5
1
F
2mg/ml data
pore-size bound
2
l
/
1
D
τ
1.5
1
F
4mg/ml data
pore-size bound
2
l
/
1
D
τ
0.5
0
0
0.5
0
0
0.5
0
0
1
1.5
0.5
dP/dF
(b)
1
1.5
0.5
dP/dF
(a)
1
1.5
0.5
dP/dF
(c)
Figure 6. Jiao and Torquato
0
τ
/
τ
3
2
1
0
0
1mg/ml data
2mg/ml data
4mg/ml data
0.5
<δ>2/τ0D1
1
1.5
Figure 7. Jiao and Torquato
Quantitative Characterization of the Microstructure and Transport Properties of Biopolymer Networks27
1
0.99
1
D
/
e
D
1mg/ml data
2mg/ml data
4mg/ml data
0.98
0
0.3
0.9
1.2
0.6
dP/dF
Figure 8. Jiao and Torquato
1
0.99
1
D
/
e
D
0.98
0.97
0
HS upper bound
dilute-limit
Ogston approx
Torquato approx
1mg/ml data
2mg/ml data
4mg/ml data
parallel cylinder data
0.01
φ
0.02
0.03
Figure 9. Jiao and Torquato
|
1608.01279 | 1 | 1608 | 2016-08-03T18:28:18 | Theory for the Acoustic Raman Modes of Proteins | [
"physics.bio-ph"
] | We present a theoretical analysis that associates the resonances of extraordinary acoustic Raman (EAR) spectroscopy [Wheaton et al., Nat Photon 9, 68 (2015)] with the collective modes of proteins. The theory uses the anisotropic elastic network model to find the protein acoustic modes, and calculates Raman intensity by treating the protein as a polarizable ellipsoid. Reasonable agreement is found between EAR spectra and our theory. Protein acoustic modes have been extensively studied theoretically to assess the role they play in protein function; this result suggests EAR as a new experimental tool for studies of protein acoustic modes. | physics.bio-ph | physics |
Theory for the Acoustic Raman Modes of Proteins
Timothy DeWolf and Reuven Gordon
Department of Electrical and Computer Engineering,
University of Victoria, British Columbia, Canada∗
(Dated: September 4, 2018)
We present a theoretical analysis that associates the resonances of extraordinary acoustic Raman
(EAR) spectroscopy [Wheaton et al., Nat Photon 9, 68 (2015)] with the collective modes of proteins.
The theory uses the anisotropic elastic network model to find the protein acoustic modes, and
calculates Raman intensity by treating the protein as a polarizable ellipsoid. Reasonable agreement
is found between EAR spectra and our theory. Protein acoustic modes have been extensively
studied theoretically to assess the role they play in protein function; this result suggests EAR as a
new experimental tool for studies of protein acoustic modes.
The central dogma of molecular biology involves one-
way information transfer from DNA to protein, a process
that directly and reliably associates a particular three-
dimensional structure with a given amino acid sequence.
Each structure is associated with a function (or func-
tions); often, for example, a particular structure cat-
alyzes a chemical reaction with remarkable selectivity.
The vibrational modes of a protein reflect its structure
and conformation, and are thought to facilitate allostery
and conformational change [1 -- 6]. Of these modes, those
with the lowest frequency are termed acoustic modes, and
represent the largest thermal fluctuations of the protein.
Whereas many spectroscopic methods can probe local-
ized resonances in a protein [7, 8], delocalized collective
modes and their role in biological function have been his-
torically difficult to measure [9]. In the gigahertz (GHz)
to low terahertz (THz) spectral window, electromagnetic
absorption experiments have to deal with high solvent
absorption and dielectric mixtures [10 -- 12]. Other experi-
mental techniques for studying acoustic protein modes in-
clude inelastic incoherent neutron scattering (IINS) [13]
and optical Kerr-effect (OKE) spectroscopy [9].
We recently reported extraordinary acoustic Raman
(EAR) spectroscopy as a way to measure resonances of
optically trapped nanoparticles [14]. In EAR, the ∼10 to
100 GHz beating of two trapping lasers creates increased
RMS fluctuation when the beat frequency matches a
Raman-active particle resonance. The frequency of vi-
brational modes in single polystyrene nanospheres were
shown to fit with Lamb's theory. The EAR spectra of
several proteins were measured, however the remaining
challenge is "...to associate the observed [protein] reso-
nances with specific motions..." [15]. Here we propose
a theory that assigns the measured EAR modes to low-
frequency Raman-active protein modes.
Our theory uses elastic network model (ENM) normal
mode analysis; elastic network models reproduce the es-
sential dynamics of low-frequency protein modes to good
accuracy [16]. We use the ENM known as the anisotropic
network model (ANM) [17, 18], as implemented in ProDy
[19].
ANM represents the potential surface of an N atom
protein (excluding hydrogen) using a network of springs
with spring constant k. ANM analyses are often done us-
ing a reduced set of atoms, namely the Cα atoms along
the protein backbone; we use an all-atom approach (ex-
cluding hydrogen) to build the elastic network. Each
spring connects a pair of atomic coordinates, but only
atoms within cutoff radius rc are connected. The matrix
of second derivatives (taken with respect to the Cartesian
coordinates of each atom) of this potential, known as the
Hessian, is then computed; it is an N × N matrix of 3× 3
(the protein coordinates are in R3) super-elements and
has the units of k. The diagonalization of this matrix
yields 3N − 6 non-zero eigenvalues λi and eigenvectors
Qi that correspond to the frequency ωi = (cid:112)λi/m and
the displacement from equilibrium of each mode i. m is
the mass of an atom; we use 13.2 amu for all atoms (a
weighted average). The six zero-valued eigenvalues corre-
spond to rotational and translational degrees of freedom.
Fig. 1(a) shows what one of these eigenvectors Qi looks
like for a protein. At this point, we have a set of mode
frequencies, as shown in Fig. 1(b).
To calculate the intensity of each mode, we need to
compute the Raman intensity of a given ANM mode from
positive and negative coordinate displacements (cid:126)ri,± =
(cid:126)r0 ± β (cid:126)Qi. The equilibrium coordinates are (cid:126)r0; (cid:126)Qi is a
unit vector in R3N and β is a small scaling parameter. We
construct a quantity closely related to the inertia tensor,
and diagonalize it to find the semi-principle axes (unit
vectors) and lengths ai,±, bi,±, ci,± of two best-fitting el-
lipsoids [20], one for each of the stretched protein coor-
dinates. Dielectric polarizability tensors αi,± for these
best-fitting ellipsoids are then computed, using analytic
expressions [21, 22] that require only the semi-principle
axis lengths and the internal and external relative dielec-
tric permittivity. We take the two permittivity values to
be i = n2 = 1.62 (protein) [23] and e = 1.332 (water).
i = (∂αi/∂Qi)0 measures
the change in polarizability due to the difference between
the protein coordinates (cid:126)ri,± [29]. We calculate this as
i ∼ αi,+ − αi,−. We also account for the possibility
α(cid:48)
that the two best-fit ellipsoids have rotated under the
action of the pair of mode displacements by rotating one
The Raman polarizability α(cid:48)
2
(a) ANM eigenmode.
(b) Frequency spectrum, plotted with
(c) Raman intensity spectrum (same
unit amplitudes.
protein modes as (b)).
FIG. 1: The method by which our protein Raman spectra are computed (shown here for carbonic anhydrase). ANM yields
a set of eigenvalues and eigenvectors; a sample eigenvector is shown in (a). The set of eigenvalues (related to the frequencies)
is shown in (b). The eigenmodes are applied to generate positive and negative displacements for each mode, and the resulting
protein shapes are fit to ellipsoids. Analytic expressions are used to compute a Raman intensity for each mode. The spectrum
shown in (c) centers a Lorentzian function, multiplied by its Raman intensity (RI), at the position of each mode shown in (b).
As seen, a small subset of the original ANM modes are dominant in the Raman spectrum.
Name
Molecular
Weight (kDa)
PDB ID ANM Spring Constant
k (kJ mol−1 A−2)
pancreatic trypsin inhibitor
carbonic anhydrase I
streptavidin
ovotransferrin
cyclooxygenase-2
6.6
29.7
52.8
76.2
274.4
5PTI [24]
1CRM [25]
3RY2 [26]
1OVT [27]
5COX [28]
1.51
1.28
1.29
0.79
1.15
TABLE I: Summary of proteins measured experimentally and the associated PDB structure files (coordinate data) used here.
of the polarizability tensors as a rank two tensor, using
the rotation matrix formed using the pair of best-fit semi-
principle axes [30]. With this, the Raman intensity Ii of
ANM mode i is given by [29]
Ii ∼ 45¯α(cid:48)2
i + 4γ(cid:48)2
(1)
i
where
¯α(cid:48)
i =
γ(cid:48)2
i =
(2)
i,yy − α(cid:48)
i,xy + α(cid:48)2
i,zz)2
i,yz + α(cid:48)2
1
3
1
2
+ (α(cid:48)
i,xx + α(cid:48)
i,xx − α(cid:48)
i,yy + α(cid:48)
i,zz)
i,yy)2 + (α(cid:48)
i,xx)2 + 6(α(cid:48)2
(α(cid:48)
((α(cid:48)
i,zz − α(cid:48)
The mean value ¯α(cid:48)
i measures change in polarizability
of a particular mode due to linear stretching; γ(cid:48)
i gives
the anisotropic contribution. Spectra (see for example
Fig. 1(c)) are constructed by centering Lorentzian func-
tions at the frequency position ωi of each mode, with
mode heights proportional to the Raman intensities Ii
(and plotting the summation of these curves). We choose
a constant Lorentzian linewidth for each spectrum.
i,zx)).
(3)
We compare our theory with previously published ex-
perimental data [14] for the five proteins listed in Table I,
and list the Research Collaboratory for Structural Bioin-
formatics Protein Data Bank (RCSB PDB) [31] struc-
tures used in computation. (The streptavidin data was
not published but was acquired during the same period.)
It is assumed that the PDB crystal coordinates are close
to the potential minimum (i.e. that the crystal coordi-
nates approximately give (cid:126)r0). By matching the atomic
mean-square fluctuations predicted by ANM (calculated
using ProDy) with the crystallographic isotropic temper-
ature factors included with the PDB crystal data, we as-
sociate a k with each protein [18]. These k values are
shown in Table I. Details of these ANM, spring constant,
ellipsoid fitting and Raman calculations are given in the
supporting information.
The ANM cutoff distance rc = 7.9 A was selected by
hand for best overall agreement between theory and ex-
periment for the five proteins. At values near this rc,
EAR mode frequencies ωi and ANM frequencies ωi are
approximately linearly proportional: ωi = ζωi. The pro-
portionality constant ζ is a free parameter in our theory;
a ζ is chosen for each protein (see supporting informa-
tion). It is the fine spectral resolution of EAR that allows
us to directly fit the spectral modes for each protein; in
prior works obtaining a Gaussian-distributed density of
states has been used as a criterion for selecting physical
values for rc [18].
ANMModes202530354045500.00.20.40.60.81.0Frequency(GHz)RI202530354045500.00.20.40.60.81.0Frequency(GHz)EXPERIMENT
THEORY
3
FIG. 2: A comparison of the experimentally measured extraordinary acoustic Raman (EAR) spectra (left) and elastic net-
work/Raman ellipsoid polarizability model spectra (right). RMS is the root-mean-squared variation in the optical trap trans-
mission. RI is the theoretical Raman intensity, calculated using an all-atom ANM with cutoff radius of rc=7.9 A. Suggested
correlations between theory and experiment of selected peaks (or groups of peaks) are letter-labelled.
TrypsinInhibitorRMSAB7080901001101201300.00.20.40.60.81.0Frequency(GHz)RIAB7080901001101201300.00.20.40.60.81.0Frequency(GHz)CarbonicAnhydraseRMSA2030405060700.00.20.40.60.81.0Frequency(GHz)RIA2030405060700.00.20.40.60.81.0Frequency(GHz)StreptavidinRMSABCDE607080901001101201300.00.20.40.60.81.0Frequency(GHz)RIABCDE607080901001101201300.00.20.40.60.81.0Frequency(GHz)ConalbuminRMSABC0204060800.00.20.40.60.81.0Frequency(GHz)RIABC0204060800.00.20.40.60.81.0Frequency(GHz)CyclooxygenaseRMSAB20304050600.00.20.40.60.81.0Frequency(GHz)RIAB20304050600.00.20.40.60.81.0Frequency(GHz)The computed spectra along with our previously ob-
tained experimental EAR spectra are shown in Fig. 2.
The visual agreement between theory (right side) and
experiment (on the left) in Fig. 2 is, in our opinion, quite
remarkable. The intensity and frequency placement of
the major peaks, as well as some of the minor peaks,
agree with the experimental data. We have made many
approximations, including the use of a "spring network"
potential in place of a more realistic potential map (e.g.
the semiempirical potentials employed in molecular dy-
namics) and the representation of protein polarizability
by the polarizability of a dielectric ellipsoid. The nor-
mal modes could also be computed in the time domain,
by combining molecular dynamics simulation with prin-
cipal component analysis to accurately capture the low-
frequency modes [32]. Wider bandwidth Raman intensity
spectra for each protein are given in the supporting in-
formation. As can be seen in the extended spectra, the
EAR data shown here have captured most of the major
Raman-active collective modes in these five proteins.
In the light of our theory, we make some comments
regarding the experiment. A past optical trapping work
by our group [33] reported on a special type of conforma-
tional change, the N-F transition, found in bovine serum
albumin (BSA); this conformation change can be viewed
as a type of denaturation as it involves a reversible un-
folding of BSA domain III [34, 35]. BSA can also be
irreversibly denatured [36]. We assume here, as we did
in the EAR experiment [14], that the trapped proteins
have not been irreversibly denatured. We also assume
that the proteins are not being reversibly unfolded or
deformed by the optical forces, so that the equilibrium
coordinates given by the x-ray crystal data will be good
approximations of the optically trapped protein coordi-
nates. Past works (e.g. [37, 38]) have suggested that an-
harmonic effects play a role in the low frequency modes
of proteins, whereas our ANM-based theory is a purely
harmonic model of protein motion.
In a Duffing oscil-
lator, for example, the absence of nonlinear effects such
as jumping, hysteresis, and bistability can be related to
the fact that harmonic driving force is sufficiently weak
[39, Chapter 7]. Thus an explanation for the seeming
unimportance of anharmonicity in our theory is that the
amplitude of the driving force is low enough that the
protein response is linear.
There has hitherto been relatively little experimental
evidence for the existence of protein collective modes --
accomplishments in protein THz spectroscopy have not
been able to conclusively connect measurements with bi-
ologically relevant collective protein motions [9]. There
has also been difficulty in assigning physical frequency
units to ENM. These results directly connect ENM mode
analysis to EAR. They provide another way to validate
ENM results, and suggest EAR as a new tool for future
experimental studies of low-frequency protein collective
modes. They also suggest that EAR may provide a way
4
to improve ENMs.
We would like to acknowledge the use of the compu-
tational resources of WestGrid (www.westgrid.ca) and
Compute Canada (www.computecanada.ca). This work
was supported in part by an NSERC Discovery Grant
and funding from the Faculty of Graduate Studies at the
University of Victoria. The protein renderings were pre-
pared using PyMOL [40] and POV-Ray [41].
∗ Electronic address: [email protected]
[1] K.-C. Chou, Biophysical Chemistry 30, 3 (1988), URL
http://www.sciencedirect.com/science/article/
pii/0301462288850026.
[2] A. Nicola, P. Delarue, and P. Senet,
in Computa-
tional Methods to Study the Structure and Dynamics
of Biomolecules and Biomolecular Processes, edited by
A. Liwo (Springer, Berlin Heidelberg, 2014).
and R. L. Jernigan, Bio-
[3] L. Yang, G. Song,
physical Journal 93,
ISSN 0006-
920
3495, URL http://www.sciencedirect.com/science/
article/pii/S0006349507713498.
(2007),
[4] S. E. Dobbins, V. I. Lesk, and M. J. E. Sternberg, Pro-
ceedings of the National Academy of Sciences 105, 10390
(2008), URL http://www.pnas.org/content/105/30/
10390.abstract.
[5] W. Zheng and D. Thirumalai, Biophysical Journal 96,
2128 (2009), URL http://www.ncbi.nlm.nih.gov/pmc/
articles/PMC2717279/.
[6] F. Tama and Y.-H. Sanejouand, Protein Engineering
14, 1 (2001), URL http://peds.oxfordjournals.org/
content/14/1/1.abstract.
[7] J. Cavanagh, W. J. Fairbrother, A. G. P. III, M. Rance,
and N. J. Skelton, Protein NMR Spectroscopy: Principles
and Practice (Academic Press, Burlington; California;
London, 2007), 2nd ed.
[8] D. Skoog, F. Holler, and S. Crouch, Principles of Instru-
mental Analysis (Thomson Brooks/Cole, 2007).
[9] D. A. Turton, H. M. Senn, T. Harwood, A. J. Lapthorn,
E. M. Ellis, and K. Wynne, Nat Commun 5 (2014), URL
http://dx.doi.org/10.1038/ncomms4999.
[10] A. Markelz, S. Whitmire, J. Hillebrecht, and R. Birge,
Physics in Medicine and Biology 47, 3797 (2002), URL
http://stacks.iop.org/0031-9155/47/i=21/a=318.
[11] J. Xu, K. W. Plaxco, and J. S. Allen, Protein Science 15,
1175 (2006), URL http://stacks.iop.org/0031-9155/
47/i=21/a=318.
[12] N. Q. Vinh, S. J. Allen, and K. W. Plaxco, Journal of
the American Chemical Society 133, 8942 (2011), URL
http://dx.doi.org/10.1021/ja200566u.
[13] H. D. Middendorf, Annual Review of Biophysics and Bio-
engineering 13, 425 (1984), URL http://dx.doi.org/
10.1146/annurev.bb.13.060184.002233.
[14] S. Wheaton, R. M. Gelfand, and R. Gordon, Nat Pho-
ton 9, 68 (2015), URL http://dx.doi.org/10.1038/
nphoton.2014.283.
[15] A. Weigel and P. Kukura, Nat Photon 9, 11 (2015), URL
http://dx.doi.org/10.1038/nphoton.2014.309.
[16] M. M. Tirion, Phys. Rev. Lett. 77, 1905 (1996), URL
http://link.aps.org/doi/10.1103/PhysRevLett.77.
1905.
Structure,
[17] P. Doruker, A. R. Atilgan, and I. Bahar, Pro-
and Bioinformat-
teins:
ics
URL
http://dx.doi.org/10.1002/1097-0134(20000815)40:
3<512::AID-PROT180>3.0.CO;2-M.
ISSN 1097-0134,
Function,
40,
512
(2000),
[18] A. Atilgan, S. Durell, R. Jernigan, M. Demirel, O. Ke-
skin, and I. Bahar, Biophysical Journal 80, 505 (2001),
ISSN 0006-3495, URL http://www.sciencedirect.com/
science/article/pii/S000634950176033X.
[19] A. Bakan, L. M. Meireles, and I. Bahar, Bioinfor-
matics 27, 1575 (2011), URL http://bioinformatics.
oxfordjournals.org/content/27/11/1575.abstract.
[20] H. Jang-Condell and L. Hernquist, The Astrophysical
Journal 548, 68 (2001), URL http://stacks.iop.org/
0004-637X/548/i=1/a=68.
[21] A. Sihvola, Electromagnetic Mixing Formulas and Appli-
cations, Electromagnetics and Radar Series (Institution
of Electrical Engineers, 1999).
[22] E. C. Stoner, The London, Edinburgh, and Dublin Philo-
sophical Magazine and Journal of Science 36, 803 (1945),
URL http://dx.doi.org/10.1080/14786444508521510.
[23] T. L. McMeekin, M. Wilensky, and M. L. Groves,
and Biophysical Research Commu-
ISSN 0006-291X, URL
Biochemical
nications 7, 151
http://www.sciencedirect.com/science/article/
pii/0006291X62901651.
(1962),
[24] A. Wlodawer, J. Walter, R. Huber, and L. Sjlin, Jour-
nal of Molecular Biology 180, 301 (1984), ISSN 0022-
2836, URL http://www.sciencedirect.com/science/
article/pii/S0022283684800066.
[25] K. K. Kannan, M. Ramanadham, and T. A. Jones, An-
nals of the New York Academy of Sciences 429, 49
(1984), ISSN 1749-6632, URL http://dx.doi.org/10.
1111/j.1749-6632.1984.tb12314.x.
[26] I. Le Trong, Z. Wang, D. E. Hyre, T. P. Lybrand, P. S.
Stayton, and R. E. Stenkamp, Acta Crystallographica
Section D 67, 813 (2011), URL http://dx.doi.org/10.
1107/S0907444911027806.
[27] H. Kurokawa, B. Mikami, and M. Hirose, Journal
(1995), ISSN 0022-
of Molecular Biology 254, 196
2836, URL http://www.sciencedirect.com/science/
article/pii/S0022283685706110.
[28] R. G. Kurumbail, A. M. Stevens, J. K. Gierse, J. J.
5
McDonald, R. A. Stegeman, J. Y. Pak, D. Gildehaus,
J. M. Miyashiro, T. D. Penning, K. Seibert, et al.,
Journal of Molecular Biology 384, 644
(1996), URL
http://dx.doi.org/10.1038/384644a0.
[29] L. Woodward, in Raman Spectroscopy: Theory and Prac-
tice, edited by H. A. Szymanski (Plenum Press, 1967).
[30] L. Hand and J. Finch, Analytical Mechanics (Cambridge
University Press, 1998).
[31] H. Berman, K. Henrick, H. Nakamura, and J. L.
Markley, Nucleic Acids Research 35, D301 (2007), URL
http://nar.oxfordjournals.org/content/35/suppl_
1/D301.abstract.
[32] I. Bahar, T. R. Lezon, L.-W. Yang, and E. Eyal, Annual
review of biophysics 39, 23 (2010), URL http://www.
ncbi.nlm.nih.gov/pmc/articles/PMC2938190/.
[33] Y. Pang and R. Gordon, Nano Letters 12, 402 (2012),
URL http://dx.doi.org/10.1021/nl203719v.
[34] V. Rosenoer, M. Oratz, and M. Rothschild, Albumin:
Structure, Function and Uses (Elsevier Science, 2014),
ISBN 9781483156880.
[35] M. Khan, Biochemical Journal 236, 307
(1986),
http://www.ncbi.nlm.nih.gov/pmc/articles/
URL
PMC1146822/.
[36] R. Wetzel, M. Becker, J. Behlke, H. Bullwitz, S. Bohm,
B. Ebert, H. Hamann, J. Krumbiegel, and G. Lassmann,
European Journal of Biochemistry 104, 469 (1980),
ISSN 1432-1033, URL http://dx.doi.org/10.1111/j.
1432-1033.1980.tb04449.x.
[37] B. Brooks and M. Karplus, Proc Natl Acad Sci USA 80,
6571 (1983), URL http://www.pnas.org/content/80/
21/6571.
[38] N. Go, T. Noguti, and T. Nishikawa, Proc Natl Acad
Sci USA 80, 3696 (1983), URL http://www.pnas.org/
content/80/12/3696.
[39] R. Enns and G. McGuire, Nonlinear Physics with Maple
for Scientists and Engineers (Birkhauser Boston, 2012),
ISBN 9781461213222.
[40] Schrodinger, LLC, The PyMOL Molecular Graphics Sys-
tem, Version 1.3, Schrodinger, LLC. (2010), [Online; ac-
cessed 2016-04-05], URL http://www.pymol.org/.
[41] Persistence of Vision Pty. Ltd., Persistence of Vision
Raytracer (Version 3.6) (2004), [Online; accessed 2016-
04-05], URL http://www.povray.org/download/.
|
1707.02382 | 4 | 1707 | 2018-07-19T06:06:11 | Nonlinear dynamics of DNA systems with inhomogeneity effects | [
"physics.bio-ph",
"cond-mat.soft",
"nlin.PS"
] | We investigate the nonlinear dynamics of the Peyrard-Bishop DNA model taking into account site dependent inhomogeneities. By means of the multiple-scale expansion in the semi-discrete approximation, the dynamics is governed by the perturbed nonlinear Schrodinger equation. We carry out a multiple-scale soliton perturbation analysis to find the effects of the variety of nonlinear inhomogeneities on the breatherlike soliton solution. During the crossing of the inhomogeneities, the coherent structure of the soliton is found stable. The global shape of the inhomogeneous molecule is merged with the shape of the homogeneous molecule. However, the velocity, the wavenumber and the angular frequency undergo a time-dependent correction that is proportional to initial width of the soliton and depends on the nature of the inhomogeneities. | physics.bio-ph | physics | Nonlinear dynamics of DNA systems with inhomogeneity effects
J. Brizar Okalya,c,∗, Alain Mvogoa,c, R. Laure Woulach´eb,c, T. Cr´epin Kofan´eb,c
aLaboratory of Biophysics, Department of Physics, Faculty of Science, University of Yaounde I, P.O. Box 812, Yaounde, Cameroon
bLaboratory of Mechanics, Department of Physics, Faculty of Science, University of Yaounde I, P.O. Box 812, Yaounde, Cameroon
cAfrican Centre of Excellence in Information and Communication Technologies, University of Yaounde I, P.O. Box 812, Yaounde,
Cameroon.
8
1
0
2
l
u
J
9
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
4
v
2
8
3
2
0
.
7
0
7
1
:
v
i
X
r
a
Abstract
We investigate the nonlinear dynamics of the Peyrard-Bishop DNA model taking into account site dependent inhomo-
geneities. By means of the multiple-scale expansion in the semi-discrete approximation, the dynamics is governed by
the perturbed nonlinear Schrodinger equation. We carry out a multiple-scale soliton perturbation analysis to find the
effects of the variety of nonlinear inhomogeneities on the breatherlike soliton solution. During the crossing of the inho-
mogeneities, the coherent structure of the soliton is found stable. The global shape of the inhomogeneous molecule is
merged with the shape of the homogeneous molecule. However, the velocity, the wavenumber and the angular frequency
undergo a time-dependent correction that is proportional to initial width of the soliton and depends on the nature of the
inhomogeneities.
Keywords: DNA; Nonlinear inhomogeneities; Breather soliton; Soliton perturbation technique.
1.
Introduction
DNA plays an important role in the carrier, protection, transmission, suppression, replication, transcription, recom-
bination, repair and mutation of genetic information in biological systems. Several processes in the cell start with the
binding of an protein enzyme at a promoter site of the DNA. This binding is known to change the conformation of DNA
by generating a nonlinear localized excitation, which causes the opening of few base pairs in the molecular chain [1]. This
excitation explains the transition conformation, the regulation of transcription, the denaturation and charge transport in
terms of polarons and bubbles [2]. Many theoretical models based on the longitudinal and transversal motions (as well
as bending, stretching and rotations) have been proposed to describe the dynamics of DNA double helix [3 -- 7]. One of
the most popular of these models is the Peyrard-Bishop (PB) model in which DNA is considered as a helicoidal structure
with a collection of particles connected with springs [8, 9].
The biological processes which are executed in nature are carried out in a setting of inhomogeneities, such as the protein
enzymes which catalyse billion chemical reactions occurring anytime in biological systems. These processes are taking
place not as isolated entities, but they do so in a molecular crowded environment. The particular biological functions
of DNA imply the presence of different sites along the strands, such as promotor (P ), coding (C), several regulatory
regions, (R1, R2, R3), terminator (T ), which contain a specific sequence of base pairs and naturally make the strands
inhomogeneous. The inhomogeneities in the DNA molecular chain can also be due to the presence of abasic site-like
nonpolar mimic of thymine or external molecules in the sequence or to the fact that the two different base pairs of the real
∗Corresponding author
Email addresses: [email protected] (J. Brizar Okaly), [email protected] (Alain Mvogo), [email protected] (R. Laure
Woulach´e), [email protected] (T. Cr´epin Kofan´e)
Preprint submitted to Elsevier
November 6, 2018
DNA, A − T and G − C, combine in different ways constituting the genetic code [10, 11]. Along the same line, in their
study on the dynamics of DNA with periodic and localized inhomogeneities both in stacking and in hydrogen bonds, using
the plane base rotator model, Daniel et al. [12, 13] have shown that the inhomogeneity is found to modulate the width
and velocity with which the open state configuration travels along the double helical chain. They also demonstrate that
the inhomogeneity introduces fluctuations in the open state configuration represented by the kink/antikink-type soliton
[12, 13]. Aguero et al. [14], using the generalized coherent states approach to averaging the quasi-spin Hamiltonian for
DNA, have found that, in a weakly saturating approximation, the formation of compacton/anticompacton pairs for the
hydrogen bond displacements is strongly influenced by the inhomogeneity due to the action of an external agent on a
specific site of the DNA [14].
The results obtained in the above studies demonstrate that, the interplay between nonlinearity and disorder in the
DNA through biological processes is yet clearly far from being understood and rather encourage us to extend the study in
another case: the case where the inhomogeneities are supposed to be due to the presence of additional molecules such as
drugs, mutants, carcinogens or other in specific sites of DNA sequences along the molecular chain and cause damages or
mutations on it. The mutations which are the accidental changes observed in the genetic code contains, occur constantly
and are due to the actions of endogenous or external agents such as redox-cycling events involving environmental toxic
agents and Fenton reactions mediated by heavy metals. Also, reactive oxygen and nitrogen compounds produced by
macrophages and neutrophils at sites of inflammation and infection arising as by-products from oxidative respiration can
lead to mutations. The above chemical agents can attack DNA, leading to adducts that impair base pairing or block
DNA replication and transcription processes, base loss, or DNA single-strand breaks (SSB). Furthermore, when the DNA-
replication apparatus encounters a SSB or certain other lesions, double-strand breaks (DSB) are formed [15 -- 18]. Cancer
usually results from a series of mutations within a single cell. Often, a faulty, damaged, or missing p53 gene are to
blame. The p53 gene provides proteins that stop mutated cells from dividing. Without this protein, cells divide unchecked
and become tumours [19]. The DNA lesions listed above can be gathered in three groups: the benefit mutations which
create genetic diversity and keep population healthy, the silent mutations which have no effect at all (such diseases do not
manifest) and the mutations which lead to diseases (base loss, SSB or DSB). Such DNA lesions are extremely toxic and
difficult to repair. Then, the most promising directions in biophysics is the study of inhomogeneous nonlinear models of
DNA, because this can give new interesting relations between the physical nonlinear properties of DNA and its biological
functioning. Studies can lead to the discovery of the new mechanisms of regulation of fundamental biological functions
of DNA, what can "bridge" the nonlinear physics of DNA and medicine [20]. For instance, breather-impurity interactions
and its scattering in the PB model was studied in detail by Kyle et al. [21]. They show that the impurity can act as a
catalyst and generates larger excitations during the fusion of the breather mode due to the nonlinearity in the system and
the one due to the mass inhomogeneity.
With this in mind, in this work, we aim to study the nonlinear dynamics of the DNA double helix taking into account
site dependence inhomogeneities, by considering the PB model. In this case, the DNA dynamics is governed by a perturbed
nonlinear Schrodinger (NLS) equation and thus, the problem boils down to solve this equation.
The paper is organized as follows. In Section 2, we propose the model Hamiltonian and derive the discrete equations
of motion for the in-phase and out-of-phase motions, respectively. Using the multiple-scale expansion in the semi-discrete
approximation, the lattice equation is reduced to the Perturbed NLS equation. In Section 3, the solitonic parameters
and the first order soliton solution are obtained through the soliton perturbation technique. Inhomogeneity effects on the
2
velocity, wavenumber, angular frequency, shape and position of the soliton solution are discussed. Section 4 concludes the
paper.
2. Model Hamiltonian of DNA dynamics and equations of motion
We consider the PB model [8] for DNA denaturation, where the degrees of freedom xn and yn associated to each base
pair correspond to the displacements of the bases from their equilibrium positions along the direction of the hydrogen
bonds that connect the two bases in a pair. A coupling between the base pairs due to the presence of the phosphate groups
along the DNA strands is assumed to be inhomogeneous [22 -- 24], so that the Hamiltonian for the model is given by:
H =Xn ( 1
2
m( x2
n + y2
n) +
1
2
Kfn[(xn − xn−1)2 + (yn − yn−1)2] + Vn(xn, yn)),
(1)
where m and K are the nucleotide mass and the elastic coupling constant in the same strand, respectively. The quantity
fn represents the inhomogeneity site dependent character introduced in the transfer of the stacking energy between nth
and (n ± 1)th base pairs. It shows that the stacking energy between neighbouring base pairs is site dependent function.
As mentioned above, the inhomogeneity represents the intercalation of the compounds (drugs, mutants or carcinogens)
between neighbouring base pairs along the DNA molecular chain without any distortion of the strands. The interactions
between two bases in a pair are done by the hydrogen bonds which are modelled by the Morse potential given by:
Vn(xn, yn) = Dhe−a(xn−yn) − 1i2
,
(2)
where D is the depth of the Morse potential, a is the width of the well. The values of parameters used to describe the
motions of the two strands are those from the dynamical and denaturation properties of DNA. They are [25, 26]: m = 300
amu, K = 0.06 eV/A2, D = 0.03 eV and a = 4.5 A−1. Our system of units (amu, A, eV) defines a time unit (t.u.) equal
to 1.018 × 10−14 seconds. The variables un and vn are introduced:
Taking into account Eqs. (2) and (3), the Hamiltonian of the system becomes
un =
xn + yn√2
and
vn =
xn − yn√2
.
H =Xn ( 1
2
m u2
n +
1
2
Kfn(un − un−1)2) +Xn ( 1
2
m v2
n +
1
2
Kfn(vn − vn−1)2 + D(cid:16)e−a√2vn − 1(cid:17)2),
(3)
(4)
where un and vn represent the in-phase and the out-of-phase motions, respectively. By using the Hamiltonian of Eq. (4),
the equation of motions of the system are given by:
mun =hKfn (un+1 − un) + Kfn−1 (un−1 − un)i,
mvn =hKfn (vn+1 − vn) + Kfn−1 (vn−1 − vn)i + 2√2aDe−a√2vn(cid:16)e−a√2vn − 1(cid:17) .
(5)
(6)
The equation of variable un(t) describes the linear waves (phonons), while the one of variable vn(t) describes the nonlinear
waves (solitons). Hence, we restrict our attention on the second equation of motion (Eq. (6)), in which the small amplitude
oscillation of the nucleotides is assumed around the bottom of the Morse potential, allowing the following transformation
[26]
vn = εψn,
3
(7)
where ε is a small parameter (ε ≪ 1). Replacing vn defined below into Eq. (6) and considering the system slightly
inhomogeneous, our investigations will be limited to the analysis of the dynamical behaviour of the stretching motion of
each base pairs, represented by the solution of Eq. (6). We assume fn = (1 + ε2gn), where gn is a site-dependent function,
which measures the inhomogeneity in the stacking. gn is treated perturbatively by assuming that, it contributes little
enough to the whole DNA dynamics which is dominated by the first and second terms of the following equation obtained
up to the third order of the Morse potential,
ψn = K1(ψn+1 − 2ψn + ψn−1) − ω2
m , α = − 3a√2
g = 4a2D
g(ψn + εαψ2
n + ε2βψ3
n) + ε2K1hgn (ψn+1 − ψn) + gn−1 (ψn−1 − ψn)i,
(8)
where K1 = K/m, ω2
3 . Equation (8) is the equation describing the dynamics of the
out-of-phase motion of a weakly inhomogeneous DNA model. It can be solved directly using the series expansion unknown
and β = 7a2
function method [27] or the exponential rational function method [28]. In this paper, we are looking for the envelope soliton
in the small amplitude approximation as a perturbed plane wave solution. For this purpose, it has been shown that the
multiple-scale expansion in the semi-discrete approximation is the most adapted technique [7, 26, 29]. This approximation
is a perturbation technique in which the amplitude is treated in the continuum limit, while the carrier waves are kept
discrete. The technique allows the study of the modulation of the wave. Thus, the soliton solution is looking for in the
form
ψn =F1,neiθn + ε(cid:0)F0,n + F2,ne2iθn(cid:1) + C.C.,
(9)
where C.C. stands for complex conjugate and θn = qnr − ωt. The quantities ω, r and q represent the optical frequency
of the linear approximation of the base pair vibrations, the distance of neighbouring bases in the same strand, and the
wavenumber, respectively. Nonlinear terms in Eq. (8) incite one to predict that, through frequency superpositions, the first
harmonics of the wave will contain terms in e±2iθn as well as terms without any exponential dependence. The amplitudes
F1,n, F0,n and F2,n will be equally considered to change slowly in space and time. For this purpose, the continuum limit
approximation and the multiple-scale expansion will be applied on those amplitudes. Thus, the amplitudes are treated
as functions of variables according to the new space and time scales zi = εiz and Ti = εit, respectively. Hence, the
solution vn(t) → v(z, t), which depends on these new sets of variables is found as a perturbation series of functions. We
εiψi(z0, z1, z2, ..., T0, T1, t2). By Taylor expansion, and up to the second order in ε, we
will consider here that v(z, t) =
∞
obtain for spatial derivatives
Pi=1
F1,n±1 =F1,±(εr)
and their temporal derivatives
∂F1
∂z1 ± (εr)2 ∂F1
∂z2
+
(εr)2
2
∂2F1
∂z2
1
+ O((εr)3),
∂F1,n
∂t
= ε
∂F1
∂T1
+ ε2 ∂F1
∂T2
+ O(ε3).
(10)
(11)
The same procedure is used for F0,n and F2,n. Using Eqs. (9), (10) and (11) together with Eq. (8) and collecting the
coefficients for the different powers of (ε, eiθn ), one obtains the angular frequency and the group velocity of the wave
and
ω2 = ω2
g + 4K1 sin2(qr/2),
vg =
.
K1r sin(qr)
ω
4
(12)
(13)
The functions F0 and F2 in Eq. (9) can be expressed through F1 as
with
F0 = −2αF12
F2 = bF 2
1 ,
b =
ω2
gα
4ω2 − ς
,
where ς = ω2
g + 4K1 sin2(qr), while F1 is a solution of the following Perturbed NLS equation
∂2F1
∂T 2
1 − 2iω
∂F1
∂T2
=2i
∂F1
∂z2
− 3ω2
K1r sin(qr) + K1r2 ∂2F1
∂z2
1
qr
2
gβF12F1 − 4K1 sin2(
cos(qr) − ω2
)g(z)F1.
gα(2F0F1 + 2F ∗1 F2)
(14)
(15)
(16)
After changing to the frame, moving at the group velocity of the carrier wave vg, by defining τ = ε2t and X = ε(z − vgt),
we get:
where
i
∂F1
∂τ
+ P
∂2F1
∂X 2 + 2QF12F1 =
2K1 sin2(qr/2)
ω
g(X + vgτ )F1,
P =
1
2ω
[K1r2 cos(qr) − v2
g ],
Q = −
ω2
g
4ωh2α(−2α + b) + 3βi.
(17)
(18)
In the following, we rescale the variable τ as τ → (τ /P ) and define a new function G such as F1(X, τ ) =p(P/Q) G(X, τ ).
Inserting the above considerations in Eq. (17), we get the Perturbed NLS equation in the form
i
∂G
∂τ
+
∂2G
∂X 2 + 2G2G = νAg(X + vgτ )G,
(19)
where ν = sin2(qr/2) ≪ 1 is assumed to be the perturbation parameter, and A = 2K1
P ω , the amplitude of the inhomogeneity.
It should be noted that Eq. (19) is the Perturbed NLS equation, regulating the dynamics of the envelope soliton
which represents the out-of-phase motion of DNA in the presence of inhomogeneity.
In the case of a system without
inhomogeneity in the lattice (g = 0) or when the wave vibrates at the frequency ω close to ωg (ω ≈ ωg), the above
equation is reduced to the following NLS equation
with the well known soliton solution [30 -- 33]
i
∂G
∂τ
+
∂2G
∂X 2 + 2G2G = 0,
G(X, τ ) =2η sechh2η(cid:16)(X − X0) − 4cτ(cid:17)iei[2c(X−X0)+4(c2−η2)τ +δ0],
(20)
(21)
where η, c, X0, δ0 are four real parameters which represent the height (as well as the width), the velocity, the initial
position and initial phase of the propagating soliton, respectively.
3. Effect of site dependence inhomogeneity on the open state
In this section, the soliton perturbation technique is used to construct the first-order perturbed soliton solution of
Eq. (19) and to find the modifications on its parameters due to the inhomogeneity. The technique allows to consider the
5
inhomogeneity as a perturbation term due to the fact that, ν is a small positive constant measuring the weakness of the
perturbation.
Following the work done by [30 -- 33], Eq. (19) is linearized by transforming the independent variable τ into several
variables
tn = νnτ,
∂τ = ∂t0 + ν∂t1 + ν2∂t2 + ...,
(22)
where the subscripts stand for partial differentiation with respect to the time tn.
It is more convenient to represent everything in the coordinate system moving with the soliton. Then, we use Z as a
new space independent variable in place of X. Under this assumption of quasi-stationary and due to the new time scale
introduced below, the soliton parameters η, c, ξ, and δ are now supposed to be function of the slow time variables t1. c
and η are independent of t0 [30 -- 33]. The one soliton (see Eq. (21)) solution of the NLS equation is rewritten here for
convenience in the form:
with
G(Z, τ, ν) = 2η sechϕeiϑ,
(23)
ϑ =
c
η
ϕ − (δ − δ0),
ϕ = 2η(Z − Z0)
Z = X − ξ,
ξt0 = −4c,
δt0 = 4(c2 + η2).
(24)
By assuming that the wave propagates at the velocity c close to the quarter of the group velocity, Eq. (24) in Eq. (19)
leads,
i
∂G
∂τ
+
∂2G
∂Z 2 + 2G2G = iR[G],
where R[G] = −iνAg(Z)G.
We assume the solution G(Z, t1, ν) to be on the form:
G(Z, t1, ν) = G0(Z, τ, ν)eiϑ,
(25)
(26)
where G0(Z, τ, ν) is the amplitude of the envelope soliton. Hence, Eq. (26) in Eq. (25) gives the linearized Perturbed NLS
equation in the form:
with
−4η2G0 +
∂2G0
∂Z 2 + 2G02G0 = νF [G0],
F [G0] =hg(Z) − 2η(Z − Z0)ct1 +(cid:16)2c
∂Z0
∂t1 −
∂δ0
∂t1(cid:17)iG0 − i
∂G0
∂t1
.
We use the Poincar´e expansion type
(27)
(28)
(29)
G0 = G(0)
0 + νG(1)
0 + ν2G(2)
0 + ...,
0 , G(1)
F [G] = F(cid:2)G(0)
0 (cid:3) + νF(cid:2)G(0)
0 (cid:3) + ...
Inserting Eq. (29) into Eq. (27) and equating the coefficients of each power of ν, we obtain the following equations at
different orders of ν:
At the order ν0, we get the following equation known as stationary NLS equation
−4η2G(0)
0 +
∂2G(0)
∂Z 2 + 2G(0)
0
0 2G(0)
0 = 0,
6
(30)
where G(0)
0 = 2η sechϕ;
At the order ν1, we get the stationary Perturbed NLS equation in the form:
−4η2G(1)
0 +
where ∗ means the complex conjugate.
∂2G(1)
0
∂Z 2 + 2G(0)2(cid:16)2G(1) + G(1)∗(cid:17) = F [G(0)
0 ],
The term G(1)
0
given in Eq. (29) is assumed to be on the form:
G(1)
0 (Z, t1, ν) = A1(Z, t1, ν) + iB1(Z, t1, ν).
Introducing Eq. (32) into Eq. (31) gives:
L1A1 ≡ −4η2A1 +
L2B1 ≡ −4η2B1 +
∂2A1
∂Z 2 + 6G(0)
∂2B1
∂Z 2 + 2G(0)
0 2A1 = ReF(cid:2)G(0)
0 (cid:3),
0 2B1 = ImF(cid:2)G(0)
0 (cid:3),
ReF(cid:2)G(0)
0 (cid:3) =hg(Z) − 2η(Z − Z0)ct1 +(cid:16)2c
∂Z0
∂t1 −
∂δ0
∂t1(cid:17)iG(0)
0 ,
ImF(cid:2)G(0)
0 (cid:3) = −
∂G(0)
0
∂t1
,
with
and
L1 = −4η2 +
L1 and L2 are two self-adjoint operators.
∂2
∂Z 2 + 6G(0)
0 ,
L2 = −4η2 +
∂2
∂Z 2 + 2(cid:12)(cid:12)G(0)
0 (cid:12)(cid:12),
(31)
(32)
(33)
(34)
(35)
3.1. Variation of the solitonic parameters
In this part, the variation of the parameters of the soliton is evaluated by assuming that, the soliton propagates with
an amplitude (as well as a width) η = η0 and a velocity c = c0, when the perturbation is switched off. To evaluate the
modifications of the solitonic parameters, it is necessary to solve the homogeneous parts of Eq. (33) which admit φ1 and
ψ2 as solutions, respectively. The non-secularity conditions [30, 32, 33] give us the following four important formulas
which determine how the solitonic parameters (i.e. the amplitude, the position, the velocity and the angular frequency)
are modified by the inhomogeneity
ηt1 =
ξt1 =
ct1 = −
1
1
1
−∞
−∞
2Z +∞
4η2 Z +∞
2Z +∞
2η Z +∞
−∞
1
RehR(cid:2)G(0)
RehR(cid:2)G(0)
ImhR(cid:2)G(0)
ImhR(cid:2)G(0)
0 (cid:3)iφ1(ϕ)dϕ,
0 (cid:3)iφ2(ϕ)dϕ,
0 (cid:3)iψ2(ϕ)dϕ,
0 (cid:3)iψ1(ϕ)dϕ,
−∞
δt1 = 2cξt1 −
with
φ1(ϕ) = sechϕ,
φ2(ϕ) = ϕ sechϕ,
where ϕ = 2η(Z − Z0). RehR[G(0)
0 ]i and ImhR[G(0)
ψ1(ϕ) = (1 − ϕ tanh ϕ) sechϕ,
0 ]i are the real and imaginary parts of R[G(0)
ψ2(ϕ) = tanh ϕ sechϕ,
0 ] given in Eq. (25).
7
(36)
(37)
(38)
(39)
(40)
To find the variation of the soliton parameters explicitly, we have to evaluate the integrals found in the right-hand
sides of Eqs. [(36)-(39)], which can be carried out only on supplying the specific forms of f (ϕ). Hence, we consider the
localized inhomogeneity in the form of the hyperbolic tangent function and the exponential inhomogeneity in the form of
exponential function separately. The results give the time dependence of the parameters of the soliton as:
ηt1 = 0,
η = η0,
ξt1 = 0,
ξτ = ξt0 ,
f (ϕ) sech2ϕ tanh ϕdϕ,
(41)
ct1 = ηAZ +∞
δt1 = AZ +∞
−∞
−∞
f (ϕ) sech2ϕ(1 − ϕ tanh ϕ)dϕ.
The first equation of the Eq. (41) shows that the amplitude (as well as the width) of the soliton remains constant, showing
that the number of base pairs which take part in the opening process remains constant during the propagation. We observe
in the second equation that the position of the soliton is not affected by the inhomogeneities. However, from the third
and fourth equations, we found that the velocity and the angular frequency of the soliton can get a correction depending
of the type of the inhomogeneities in the lattice.
The localized inhomogeneity can represent the intercalation of a compound between neighbouring base pairs or the
presence of a defect on an abasic site in the DNA chain. In order to understand the effects of this type of inhomogeneity
in the lattice, we substitute f (ϕ) = tanh(ϕ), in Eq. (41), and after integration, we obtain ct1 = 2
3 ηA and δt1 = 0, which
can be written in terms of the original time variable τ by using the expression cτ = ct0 + νct1 and δτ = δt0 + νδt1 as:
c = c0 +
2
3
(νA)ητ,
Ων ≡ δτ = 4(c2
0 + η2) +
8
3
(νA)ηh2c0τ + (νA)ητ 2i.
(42)
Equation (42) demonstrates that, the localized inhomogeneity in the above form affects the velocity and the angular
frequency of the soliton. They get a correction during the crossing of this inhomogeneity. The nature of the correction
depends of the nature of the inhomogeneity represented by the sign of its amplitude A which can be either positive or
negative. When A is positive (A > 0), the inhomogeneity corresponds to an energetic barrier and on the other hand, when
A is negative (A < 0), the inhomogeneity behaves as a potential well [13].
When A is greater than zero, the correction is positive. That makes the velocity and the angular frequency increasing
during the crossing of the inhomogeneity. The amount of these increments is proportional to the initial height (as
well as the initial width) of the soliton. The higher the initial amplitude of the soliton, the greater the increments.
The increasing in the velocity of the soliton helps to overcome the barrier due to the inhomogeneity and the soliton
will propagate easily along the inhomogeneous DNA chain without formation of a bound state. Reporting these
observations in Eq. (23), we notice that during the crossing of the inhomogeneity, the soliton propagates and vibrates
quickly, and the DNA breathing mode becomes faster.
In the case of A less than zero, the correction is negative. That makes the frequency and the velocity decreasing. The
decreasing of the velocity slows down the soliton. The soliton stops and vanishes when the original time satisfies the
condition τ = t0 = −c0
Bi(νA)η , where Bi is a constant which depends on the inhomogeneity. The propagating time of
the soliton depends on its initial velocity and width. The wider the soliton, the greater the propagating time.
When A = 0, the velocity and the angular frequency of the soliton remain constant.
8
In all the above cases, the soliton which is a coherent structure formed by involving a few base pairs, moves along the
helical chain in the form of bubble without dissipation or any other form of deformation is found stable.
Next, we consider f (ϕ) = eϕ. This case corresponds to the physically interesting problem of the exponential distribution
of similar molecules along the DNA helical chain. We have ct1 = π
2 ηA and δt1 = 0. Written in the original time variable,
we get
c = c0 +
π
2
(νA)ητ,
Ων ≡ δτ = 4(c2
0 + η2) + 2π(νA)ηh2c0τ +
π
2
(νη)Aτ 2i.
(43)
The same observations as in the previous type of inhomogeneity are also done here. We notice from the results that
considering the localized and exponential inhomogeneities, the propagating velocity and frequency of the soliton undergo
a correction during the crossing of the inhomogeneities. In Figures (1) and (2), the time-evolution of the velocity and
angular frequency of the soliton are depicted as functions of the type and nature of the inhomogeneities in the lattice.
The figures show that the correction terms can be positive or negative according to the nature of the inhomogeneities.
When comparing Eq. (42) and Eq. (43), we notice that, the absolute value of the correction terms is greater in the
case of exponential inhomogeneity as can be seen in the above Figures. This is because in this case, the inhomogeneities
occur exponentially in the entire chain of the DNA molecule in terms of external agents.
3.2. First-order perturbed soliton solution
In this part of our work, we will seek for the breatherlike soliton solution for Eq. (25). As we know, once the seed
solution is chosen as breather soliton solution via Eq. (23), the first-order perturbed breatherlike soliton solution can be
constructed using the soliton perturbation technique by solving Eq. (33). Then, we report the solutions A1 and B1 in
Eq. (32). We are solving first the homogeneous part of the first equation in Eq. (33), which admits the following two
particular solutions:
and
A12 =
A11 = sechϕ tanh ϕ,
ϕ sechϕ tanh ϕ +
1
2
tanh ϕ sinh ϕ − sechϕi.
1
2ηh 3
2
Then, the general solution can be obtained by using the following formula:
A1 = C1A11 + C2A12 −
1
2η
A11Z ϕ
−∞
A12ReF(cid:2)G(0)
0 (cid:3)dϕ +
1
2η
A12Z ϕ
−∞
A11ReF(cid:2)G(0)
0 (cid:3)dϕ,
(44)
(45)
(46)
where C1 and C2 are two arbitrary constants to be determined. We construct the solution A1, by substituting the
expressions of A11, A12 and ReF (G(0)
0 ) in Eq. (46). After the evaluation of the integrals, we obtain:
A1 = −
+
+
1
∂δ0
∂Z0
∂t1 −
1
1
1
2(cid:16)2c
∂t1(cid:17)i sechϕ +hC1 +
2ηhC2 −
∂t1(cid:17)i sechϕ tanh ϕ
8η2 ct1 (1 − tanh ϕ)h1 + ϕe2ϕ(1 − tanh ϕ)ih3ϕ sechϕ tanh ϕ + tanh ϕ sinh ϕ − 2 sechϕi
8η2 ct1 ϕ2 sechϕ tanh ϕh1 − 3e2ϕ(1 − tanh ϕ)i +
sinh ϕ tanh ϕ − I + J,
ϕ(cid:16)2c
ϕ −
C2
4η
∂δ0
Z0
∂t1 −
3C2
4η
1
4η
where I and J are the integrals depending on the form of the inhomogeneities and are given by:
I =
1
2η
AA11Z ϕ
−∞
A12f (ϕ)G(0)
0 dϕ,
J =
1
2η
AA12Z ϕ
−∞
A11f (ϕ)G(0)
0 dϕ
9
(47)
(48)
The term in C2
4η sinh ϕ tanh ϕ is a secular term which makes the solution unbounded. The above term can be removed if
we assume that the arbitrary constant C2 = 0. Using the boundary conditions A1ϕ=0 = µ and A1ϕϕ=0 = 0, we obtain
4η(cid:16)2c ∂Z0
8η2 ct1 and C1 = 0. By using the above results in Eq. (47), we get the general solution in the form:
∂t1 − ∂δ0
1
∂t1(cid:17) = µ + 1
A1 =(cid:16)µ +
1
1
1
+
8η2 ct1(cid:17)h1 − ϕ tanh ϕi sechϕ +
8η2 ct1(1 − tanh ϕ)h1 + ϕe2ϕ(1 − tanh ϕ)ih3ϕ sechϕ tanh ϕ + tanh ϕ sinh ϕ − 2 sechϕi
− I + J.
8η2 ct1 ϕ2 sechϕ tanh ϕh1 − 3e2ϕ(1 − tanh ϕ)i
(49)
Now, we give the solution of the second equation in Eq. (33). Its homogeneous part admits the particular solutions given
by:
B11 = sechϕ, B12 =
1
4η
(ϕ sechϕ + sinh ϕ),
while the general form of B1 is given by:
B1 = C3B11 + C4B12 −
1
2η
B11Z ϕ
−∞
B12ImF (G(0)
0 )dϕ +
1
2η
B12Z ϕ
−∞
B11ImF (G(0)
0 )dϕ,
(50)
(51)
where C3 and C4 are two arbitrary constants to be determined. We follow the same procedure by removing the secular
terms (C3 = C4 = 0). We get also ∂Z0
= 0. Applying the following boundary conditions B1ϕ=0 = B1ϕϕ=0 = 0, we have
∂t1
Taking into account the above results, the general form of the envelope soliton is
B1 = 0.
G0(ϕ, τ, ν) =2η sechϕ + ν((cid:16)µ +
1
8η2 ct1(cid:17)h1 − ϕ tanh ϕi sechϕ +
8η2 ct1ϕ2 sechϕ tanh ϕh1 − 3e2ϕ(1 − tanh ϕ)i
8η2 ct1(1 − tanh ϕ)h1 + ϕe2ϕ(1 − tanh ϕ)ih3ϕ sechϕ tanh ϕ + tanh ϕ sinh ϕ − 2 sechϕi − I + J),
1
+
1
(52)
(53)
where I and J depend on the type of inhomogeneity and are given in Eq. (48). The effects of the inhomogeneities on the
shape of the first order perturbed soliton is therefore reduced to the integration of I and J.
We construct the first order perturbed soliton solution by substituting the corresponding values of ct1 , δt1 for each
expression of f (ϕ) given below, and evaluate the integrals I and J, which involve very lengthy algebra. Thus, in the case
f (ϕ) = tanh(ϕ), we use the values of ct1 and δt1 of Eq. (42). Eq. (53) becomes,
G0 =2η sechϕ + ν((cid:16)µ +
ηA
ηA
12η2(cid:17)(1 − ϕ tanh ϕ) sechϕ +
ηA
12η2 ϕ2 sechϕ tanh ϕh1 − 3e2ϕ(1 − tanh ϕ)i
+
−
+
1
2
1
2η
12η2 (1 − tanh ϕ)h1 + ϕe2ϕ(1 − tanh ϕ)ih3ϕ sechϕ tanh ϕ + tanh ϕ sinh ϕ − 2 sechϕi
A sechϕ tanh ϕhϕ −
sech2ϕi +
tanh2 ϕ sinh ϕ − sechϕ tanh ϕi).
(1 − tanh ϕ)3(ϕ + e2ϕ + 3ϕe4ϕ + e4ϕ) −
A(1 − sech2ϕ)h 3
ln(e2ϕ + 1) −
ϕ sechϕ tanh2 ϕ +
1
6η
1
4
1
2
1
8
1
2
2
tanh2 ϕ
(54)
We then repeat the same procedure for constructing the first order perturbed soliton solution in the case of f (ϕ) = eϕ
10
and we obtain the perturbed soliton solution as:
G0 =2η sechϕ + ν((cid:16)µ +
πηA
πηA
16η2(cid:17)(1 − ϕ tanh ϕ) sechϕ +
πηA
16η2 ϕ2 sechϕ tanh ϕh1 − 3e2ϕ(1 − tanh ϕ)i
+
+
−
1
4η
1
4η
16η2 (1 − tanh ϕ)h1 + ϕe2ϕ(1 − tanh ϕ)ih3ϕ sechϕ tanh ϕ + tanh ϕ sinh ϕ − 2 sechϕi
A sechϕ tanh ϕh 3
A(cid:16)3ϕ sechϕ tanh ϕ + tanh ϕ sinh ϕ − 2 sechϕ(cid:17)(cid:16) 1
eϕ(1 − tanh ϕ)2(ϕ + 3ϕe2ϕ + e2ϕ + 1) + 3 tanh sinh ϕ − 4 cosh ϕ − sinh ϕi
tanh ϕ sechϕ − tanh ϕ sinh ϕ + cosh ϕ − arctan(eϕ)(cid:17)).
4
2
(55)
Equations (54) and (55) are the breatherlike solitons which represent the envelope of the soliton solutions. They describe
the open state configurations in the individual strand of the DNA, which collectively represents bubbles moving along the
inhomogeneous DNA molecule.
We have depicted in Figure 3, the 3D schematic representation of the square of the envelope solitons as functions of
the type of inhomogeneity in the lattice. Figure 3a presents the breather soliton moving in the homogeneous DNA chain.
The case of localized inhomogeneity is depicted in Figure 3b, while the case of exponential inhomogeneity is depicted in
Figure 3c. We notice that, in both inhomogeneous cases, the robust nature of the breatherlike soliton is not modified as
it propagates along the DNA molecule. The global shape of the molecule with inhomogeneities is merged with its shape
without taking into account the inhomogeneities. Similar results were observed by Mvogo et al. [34] in their investigations
on the effects of localized inhomogeneity in the form of hyperbolic secant function in the α-helical proteins chain. Also we
found in Ref. [35] that, using the numerical methods, the authors demonstrate that neither the opening of base pairs in
DNA molecule nor the topological character of the breatherlike soliton are affected by the localized inhomogeneity in the
form of hyperbolic secant function introduced in stacking and hydrogen bonding energies of DNA.
Considering Eq. (7) together with Eq. (9), we assume that the configuration of the molecule can be fully described by
the following soliton solution with a rescaled time
where
vn =εG0 cos(Rnr − Ωt),
R = q + 2εc, Ω = ω + 2εcvg + Ων .
(56)
(57)
R and Ω are the "effective "wavenumber and angular frequency of the soliton solution, respectively. G0, c and Ων represent
the envelope soliton, its velocity and angular frequency, respectively. They depend on the type of inhomogeneity and are
given by [Eq. (42) and Eq. (54)] and [Eq. (43) and Eq. (55)] for the localized and the exponential inhomogeneities,
respectively. Eq. (56) represents the stretching of the base pairs also known as the breathing modes experimentally
observed in DNA molecule [36, 37]. Since the wave velocity c depends on the type of inhomogeneity in the lattice, from
the first equation of Eq. (57), we notice that the wavenumber R is a function of the inhomogeneities. It changes as time
progressed due to the inhomogeneities.
The configuration of the molecule and the elongation of the out-of-phase motion are depicted as functions of the time,
depending on the type and nature of the inhomogeneities present in the lattice. Figures 4 and 5 represent the configuration
of the molecule, while Figures 6 and 7 present the elongation of the out-of-phase motion. As predicted by Eqs. (42),
(43) and (57), we observe that the wave velocity, the wavenumber and the angular frequency of the soliton solution get a
correction which can be positive or negative depending on the nature of the inhomogeneities:
11
For A > 0, the inhomogeneity behaves as an energetic barrier. The above parameters increase as time increases due to
the inhomogeneities. Figure 4 shows the increasing of the wavenumber and wave velocity (Figure 4d), while Figure 6
shows the increasing of the angular frequency. The increasing of the velocity helps the soliton to overcome this
energetic barrier. Therefore, the soliton can propagate easily without formation of a bound state.
For A < 0, the decreasing of the above wave parameters is observed in Figure 5 for the wavenumber, and in Figure 7 for
the angular frequency. In this case, the inhomogeneity behaves as a potential well in which the soliton is trapped.
Thus, the soliton slows down, stops and vanishes at t0 = −c0
Bi(νA)η , where Bi is equal to 2/3 and π/2 for localized and
exponential inhomogeneities, respectively.
For A = 0, the inhomogeneity is switched off. This case corresponds to the soliton travelling in a homogeneous media
with the constant velocity, wavenumber and angular frequency.
Also, as in Figures 1 and 2, we observe that the absolute value of the corrections in the velocity and wavenumber (see
Figure 4 and Figure 5), and in the angular frequency (see Figure 6 and Figure 7) of the soliton is greater in the case of
exponential inhomogeneity. We also notice that the inhomogeneities do not affect the bubble sizes. The height and the
width of the bubble remain constant during its propagation in the inhomogeneous DNA molecule.
It is clearly known that the transcription process is a very biological complex phenomenon [38]. Thus, the increasing
of the velocity, the wavenumber and the angular frequency of the soliton during the crossing of the inhomogeneities makes
difficult the execution of the biological functions of DNA such as reading, transcription and recombination of the genetic
code to mention a few. These difficulties undoubtedly involve errors which can rise to mutations, damaging, missing of
genes and biological diseases such as: sickle cell anemia, heart diseases, high blood pressure, Alzheimer's disease, diabetes,
cancer, obesity, eye disease, epilepsy or stroke-like episodes which are due to mutations or disorder caused by a combination
of environmental factors and mutations.
4. Conclusion
The wave dynamics of the PB model of DNA at the physiological temperature was studied. The classical PB model
assumes the bases to be equal. In our model, we include the differences of bases and sequences through site dependent
spring-constant. Using the multiple-scale expansion method in the semi-discrete approximation, the out-of-phase motion
has been described by the Perturbed NLS equation. In the homogeneous limit, the dynamics is governed by the breather
soliton of the integrable NLS equation. Indeed, to understand the effect of inhomogeneities on the base pairs opening, we
carried out a perturbation analysis using multiple-scale soliton perturbation technique. For implement this, we linearized
the Perturbed NLS equation with the aids of Poincar´e-type asymptotic expansion. The breatherlike soliton solution
which represents the opening of base pairs travelling along the inhomogeneous DNA chain in the form of bubble has been
constructed for different forms of the inhomogeneities (localized and exponential).
The results showed that the bubble profile is not affected by the inhomogeneities. However, the velocity with which the
base pairs are opening or closing, the wavenumber with which the soliton propagates and angular frequency with which
the base pairs vibrate increase, decrease or remain uniform and even the soliton stops depending on the nature of the
inhomogeneities. The high velocity, wavenumber and angular frequency of the soliton can create disorder in the execution
of DNA biological functions and make coding or reading errors which can leads to varieties of biological diseases.
12
The inhomogeneous DNA models can explain DNA biological functions such as replication, transcription and recom-
bination more viably than the homogeneous one, since it can explain and predict the biological diseases due to genetic
mutations. The number of DNA damages in a single human cell exceeds 10, 000 every day [39], and must be counteracted
by special DNA repair processes. Long-range interactions (LRI) are ubiquitous in DNA molecule, they play a crucial role
in the stabilization of the molecule [25]. Hence, it is important to understand the interplay between inhomogeneity and
LRI [40]. These studies are now in progress. Despite the relevance of this work, the impact of inhomogeneities on the
DNA biological processes is not yet clearly understood, since the nature generally selects inhomogeneous DNA and the
gap between the nonlinear physics of DNA and medicine is still big.
Acknowledgements
J. B. Okaly is in debt to Dr. Ndzana Fabien II of the University of Maroua, Maroua- Cameroon for some recommendations
and fruitful discussions.
References
References
[1] L. Styer, Biochemistry, fourth ed., W. H. Freeman and Company, New York, 1995.
[2] G. Kalosakas, K. Q. Rasmussen and A. R. Bishop, Nonlinear excitations in DNA: polarons and bubbles, Synth.
Met. 141 (2004) 93-97.
[3] S. W. Englander, N. R. Kallenbach, A. J. Heeger, J. A. Krumhansl and S. Litwint, Nature of the Open State
in Long Polynucleotide Double Helices: Possibility of Soliton Excitations, Proc. Natl. Acad. Sci. 77 (1980)
7222-7226.
[4] L. V. Yakushevich, Nonlinear DNA dynamics: hierarchy of the models, Physica D 79 (1994) 77-86.
[5] L. V. Yakushevich, Nonlinear DNA dynamics: a new model, Phys. Lett. A 136 (1989) 413-417.
[6] K. De-Xing, L. Sen-Yue and Z. Jin, Nonlinear dynamics in a new double chain-model of DNA, Commun. Theor.
Phys. 36 (2001) 737-742.
[7] J. B. Okaly, A. Mvogo, R. L. Woulach´e and T. C. Kofan´e, Semi-discrete Breather in a Helicoidal DNA Double
Chain-Model (Unpublished results).
[8] M. Peyrard and A. R. Bishop, Statistical Mechanics of a Nonlinear Model for DNA Denaturation, Phys. Rev.
Lett. 62 (1989) 2755-2758.
[9] T. Dauxois, Dynamics of breather modes in a nonlinear "helicoidal" model of DNA, Phys. Lett. A. 159 (1991)
390-395.
[10] J. Ladik, J. Cizek, Probable physical mechanisms of the activation of oncogenes through carcinogens, Int. J.
Quantum Chem. 26 (1984) 955-964.
13
[11] 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 Oligonucleotide,
J. Am. Chem. Soc. 121 (1999) 8653-8654.
[12] M. Daniel and V. Vasumathi, Perturbed soliton excitations in the DNA double helix, Physica D 231 (2007)
10-29.
[13] M. Daniel and V. Vasumathi, Nonlinear molecular excitations in a completely inhomogeneous DNA chain, Phys.
Lett. A 372 (2008) 5144-5151.
[14] M. A. Aguero , T. L. Belyaeva and V. N. Serkin, Compacton anti-compacton pair for hydrogen bonds and
rotational waves in DNA dynamics, Commun. Nonlinear Sci. Numer. Simulat. 16 (2011) 3071-3080.
[15] M. Valko, C. J. Rhodes, J. Moncol, M. Izakovic and M. Mazur, Free radicals, metals and antioxidants in
oxidative stress-induced cancer, Chem. Biol. Interact. 160 (2006) 1-40.
[16] S. Kawanishi, Y. Hiraku, S. Pinlaor and N. Ma, Oxidative and nitrative DNA damage in animals and patients
with inflammatory diseases in relation to inflammation-related carcinogenesis, Biol. Chem. 387 (2006) 365-372.
[17] K. K. Khanna, S. P. Jackson, DNA double-strand breaks: signaling, repair and the cancer connection, Nature
Genet. 27 (2001) 247-254.
[18] S. P. Jackson and R. J. Bartek, The DNA-damage response in human biology and disease, Nature Rev. 461
(2009) 1071-1078.
[19] J. H. Bielas, K. R. Loeb, B. P. Rubin, L. D. True and Loeb, Human cancers express a mutator phenotype, Proc.
Natl Acad. Sci. 103 (2006) 18238-18242.
[20] L. V. Yakushevich, Nonlinear Physics of DNA, Wiley, Chichester, 2004.
[21] K. Forinash, M. Peyrard and B. Malomed, Interaction of discrete breathers with impurity modes, Phys. Rev.
E 49 (1994) 3400-3411.
[22] W. Gratzer Association of nucleic-acid bases in aqueous solution: A solvent partition study. Eur. J. Biochem.
10 (1969) 184-187.
[23] D. B. Davies Co-operative conformational properties of nucleosides, nucleotides and nucleotidyl units in solu-
tion, B. Pullmlln (ed.), Nuclear Magnetic Resonance Spectroscopy in Molecular Biology, D. Reidel Publishing
Company, Dordrecht, Holland, 1978, pp. 71-85.
[24] R. L. Ornstein, R. Rein, D. L. Breen and R. D. MacElroy An optimized potential function for calculation of
nucleic acid interaction energies. I. Base stacking. Biopolymers 17 (1978) 2341-2360.
[25] J. B. Okaly, A. Mvogo, R. L. Woulach´e and T. C. Kofan´e, Nonlinear dynamics of damped DNA systems with
long-range interactions, Commun. Nonlinear Sci. Numer. Simulat., 55 (2018) 183-193.
[26] M. Peyrard, Nonlinear dynamics and statistical physics of DNA, Nonlinearity 17 (2004) R1-R40.
14
[27] S. Zdravkovi´c and S. Zekovi´c, Nonlinear dynamics of microtubules and series expansion unknown function
method, Chinese. J. Phys. 55 (2017) 2400-2406.
[28] E. Tala-Tebue, Z. I. Djoufack, D .C. Tsobgni-Fozap, A. Kenfack-Jiotsa, F. Kapche-Tagne, T. C. Kofan´e,
Traveling wave solutions along microtubules and in the Zhiber-Shabat equation, Chinese J. Phys., 55 (2017)
939-946.
[29] M. Remoissenet, Low-amplitude breather and envelope solitons in quasi-one-dimensional physical models, Phys.
Rev. B 33 (1986) 2386-2392.
[30] J. Yan, Y. Tang, and G. Zhou, Direct approach to the study of soliton perturbations of the nonlinear Schrodinger
equation and the sine-Gordon equation, Phys. Rev. E 58 (1998) 1064-1073.
[31] V. Vasumathi and M. Daniel, Base-pair opening and bubble transport in a DNA double helix induced by a
protein molecule in a viscous medium, Phys. Rev. E 80 (1-9) (2009), 061904.
[32] V. I. Karpman, Soliton evolution in the presence of perturbation, Phys. Scr. 20 (1979) 462-478.
[33] V. I. Karpman and V. V. Solov'ev, A perturbation theory for soliton systems, Physica D 3 (1981) 142-164.
[34] A. Mvogo, G.H. Ben-Bolie and T. C. Kofan´e, Solitary waves in an inhomogeneous chain of α-helical proteins,
Int. J. of Mod. Phys B 28 (1-14) (2014), 1450109.
[35] M. Saha and T. C. Kofan´e, inhomogeneities and nonlinear dynamics of a helical DNA interacting with a
RNA-polymerase, Phys. Scr. 89 (1-7) (2014), 085003.
[36] B. F. Putnam, L. L. Van Zandt, E. W. Prohofsky, M. N. Mei, Resonant and localized breathing modes in
terminal regions of the DNA double helix, Biophys J. 35 (1981) 271-287.
[37] E. W. Prohofsky, K. C. Lu, L. L. Van Zandt and B.F. Putnam, Breathing modes and induced resonant melting
of the double helix, Phys. Lett. A 70 (1979) 492-494.
[38] X. Yang, Y. Wu a, Z. Yuan, Characteristics of mRNA dynamics in a multi-on model of stochastic transcription
with regulation, Chinese J. Phys. 55 (2017) 508-518.
[39] A. Klungland, Y-G Yang, Endogenous DNA Damage and Repair Enzymes: -A short summary of the scientific
achievements of Tomas Lindahl, Nobel Laureate in Chemistry 2015, Genomics Proteomics Bioinformatics 14
(2016) 122-125.
[40] M. Saha and T. C. Kofan´e, Long-range interactions between adjacent and distant bases in a DNA and their
impact on the ribonucleic acid polymerase-DNA dynamics, Chaos 22 (1-12) (2012), 013116.
15
0.7
0.6
0.5
c
0.4
0.3
0.2
0.1
0
(a): A=1
Localized inhomogeneity
Exponential inhomogeneity
(b): A=-1
Localized inhomogeneity
Exponential inhomogeneity
0.2
0.18
0.16
0.14
0.12
c
0.1
0.08
0.06
0.04
0.02
0.5
1
1.5
τ
2
2.5
3
0
0.5
1
1.5
τ
2
2.5
3
Figure 1: Variation of the velocity of the soliton as a function of the time, depending on the nature of the inhomogeneity in the lattice for
ν = 0.1, c0 = 0.18, and η = 1.
4.9
4.8
4.7
4.6
ν
Ω
4.5
4.4
4.3
4.2
4.1
0
(a): A=1
Localized inhomogeneitiy
Exponential inhomogeneity
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
ν
Ω
4.14
4.12
4.1
4.08
4.06
4.04
4.02
4
0
(b): A=-1
Localized inhomogeneity
Exponential inhomogeneity
0.2
0.4
0.6
0.8
1
1.2
1.4
1.6
1.8
τ
τ
Figure 2: Variation of the angular frequency of the soliton as a function of the time, depending on the nature of the inhomogeneity in the
lattice. The parameters are the same as in Figure 1.
16
(a): Homogeneous case ( ν=0.00)
(b): Localized ihnomogeneity ( ν=0.01)
2
0
G
0.04
0.03
0.02
0.01
0
0
10
20
30
X
40
50
0
1
2
τ
5
4
3
2
0
G
0.04
0.03
0.02
0.01
0
0
10
20
30
X
40
50
0
1
2
τ
5
4
3
(c): Exponential ihnomogeneity ( ν=0.01)
2
0
G
0.04
0.03
0.02
0.01
0
0
10
20
30
X
40
50
0
1
2
τ
5
4
3
Figure 3: Dynamical evolutions of the first-order perturbed breatherlike solitons [see (54)-(55)], with the parameters: ε = 0.5, c0 = 0.01,
η = 0.1, ν = 0.01, µ = 0, Z0 = 0 and A = 1.
17
(a): Homogeneous case ( ν=0.0)
2
4
6
10
8
12
Base pairs (n)
14
16
18
20
(c): Exponential inhomogeneity ( ν=0.5)
0.2
0.15
0.1
0.05
n
v
0
-0.05
-0.1
-0.15
-0.2
0.2
(b): Localized inhomogeneity( ν=0.5)
14
16
18
20
22
24
26
28
30
32
Base pairs (n)
(d)
0.2
0.15
0.1
0.05
n
v
0
-0.05
-0.1
-0.15
-0.2
0
0.2
0.15
0.1
0.05
n
v
0
-0.05
-0.1
-0.15
0.15
1
2
3
0.1
0.05
n
v
0
-0.05
-0.1
-0.15
-0.2
0
5
10
15
25
20
30
Base pairs (n)
35
40
45
50
-0.2
30
32
34
36
40
38
42
Base pairs (n)
44
46
48
50
Figure 4: Stretching of the nucleotide pair vs base pairs at t = 18 t.u., depending on the type of the inhomogeneity in the lattice, for ε = 0.1
c0 = 0.18, η = 1, q = π
8 and A = 1. The red lines (bubble (1)) represent the bubble propagating in the homogeneous DNA chain, while the
blue lines and the green lines (bubble (2) and bubble (3)) represent the bubbles propagating in the DNA chain in the presence of localized and
exponential inhomogeneities, respectively. Dashed lines represent the envelope, while the solid lines represent the stretching of the base pairs
18
0.2
0.15
0.1
0.05
n
v
0
-0.05
-0.1
-0.15
-0.2
0
(a): Homogeneous case ( ν=0.0)
2
4
6
10
8
12
Base pairs (n)
14
16
18
20
0.2
0.15
0.1
0.05
n
v
0
-0.05
-0.1
-0.15
-0.2
0
(b): Localized inhomogeneity ( ν=0.9)
2
4
6
10
8
12
Base pairs (n)
14
16
18
20
0.2
0.15
0.1
0.05
n
v
0
-0.05
-0.1
-0.15
-0.2
0
(c): Exponential inhomogeneity ( ν=0.9)
2
4
6
10
8
12
Base pairs (n)
14
16
18
20
Figure 5: Stretching of the nucleotide pair vs base pairs at t = 0.9 t.u., depending on the type of the inhomogeneity in the lattice, for A = −1.
The other parameters are the same as in Figure 4. .
19
0.4
0.3
0.2
0.1
n
v
0
-0.1
-0.2
-0.3
-0.4
0
(a): Homogeneous case ( ν=0.0)
0.1
0.2
0.3
0.4
0.5
Time (t.u.)
0.6
0.7
0.8
0.9
0.4
0.3
0.2
0.1
n
v
0
-0.1
-0.2
-0.3
-0.4
0
(b): Localized inhomogeneity ( ν=0.1)
0.1
0.2
0.3
0.4
0.5
Time (t.u.)
0.6
0.7
0.8
0.9
0.4
0.3
0.2
0.1
n
v
0
-0.1
-0.2
-0.3
-0.4
0
(c): Exponential inhomogeneity ( ν=0.1)
0.1
0.2
0.3
0.4
0.5
Time (t.u.)
0.6
0.7
0.8
0.9
Figure 6: Elongation of the out-of-phase motion vs time, depending on the type of the inhomogeneity in the lattice, for ε = 0.2 c0 = 50, η = 1,
q = π
8 , n = 0 and A = 1. Panel (a) represents the homogeneous case, panel (b) represents the case of localized inhomogeneity, while panel (c)
represents the case of exponential inhomogeneity. Dashed lines represent the envelope, while the solid lines represent the stretching of the base
pairs.
20
0.4
0.3
0.2
0.1
n
v
0
-0.1
-0.2
-0.3
-0.4
0
(a): Homogeneous case ( ν=0.00)
0.1
0.2
0.3
0.4
0.5
Time (t.u.)
0.6
0.4
0.8
0.9
0.4
0.3
0.2
0.1
n
v
0
-0.1
-0.2
-0.3
-0.4
0
(b): Localized inhomogeneity ( ν=0.04)
0.1
0.2
0.3
0.4
0.5
Time (t.u.)
0.6
0.7
0.8
0.9
0.4
0.3
0.2
0.1
n
v
0
-0.1
-0.2
-0.3
-0.4
0
(c): Exponential inhomogeneity ( ν=0.04)
0.1
0.2
0.3
0.4
0.5
Time (t.u.)
0.6
0.7
0.8
0.9
Figure 7: Elongation of the out-of-phase motion vs time, depending on the type of the inhomogeneity in the lattice for A = −1. The other
parameters are the same as in Figure 6.
21
|
1207.5202 | 2 | 1207 | 2012-09-14T07:46:53 | Temperature and force dependence of nanoscale electron transport via the Cu protein Azurin | [
"physics.bio-ph",
"cond-mat.mtrl-sci",
"q-bio.BM"
] | The mechanisms of solid-state electron transport (ETp) via a monolayer of immobilized Azurin (Az) was examined by conducting probe atomic force microscopy (CP-AFM), both as function of temperature (248 - 373K) and of applied tip force (6-12 nN). By varying both temperature and force in CP-AFM, we find that the ETp mechanism can alter with a change in the force applied via the tip to the proteins. As the applied force increases, ETp via Az changes from temperature-independent to thermally activated at high temperatures. This is in contrast to the Cu-depleted form of Az (apo-Az), where increasing the applied force causes only small quantitative effects, that fit with a decrease in electrode spacing. At low force ETp via holo-Az is temperature-independent and thermally activated via apo-Az. This observation agrees with macroscopic-scale measurements, thus confirming that the difference in ETp dependence on temperature between holo- and apo-Az is an inherent one that may reflect a difference in rigidity between the two forms. An important implication of these results, which depend on CP-AFM measurements over a significant temperature range, is that for ETp measurements on floppy systems, such as proteins, the stress applied to the sample should be kept constant or, at least controlled during measurement. | physics.bio-ph | physics | Temperature and force dependence of nanoscale
electron transport via the Cu protein Azurin
Wenjie Li,† Lior Sepunaru,†, ‡ Nadav Amdursky,†,‡ Sidney R. Cohen&, Israel Pecht,§ Mordechai
Sheves,*,‡ and David Cahen*,†
†Department of Materials & Interfaces, ‡Department of Organic Chemistry, &Chemical Research
Support and §Department of Immunology, Weizmann Institute of Science, Rehovot 76100, Israel
*Corresponding authors:
David Cahen
e-mail: [email protected]
Mordechai Sheves
e-mail: [email protected]
TOC graphic:
1
Abstract
The mechanisms of solid-state electron transport (ETp) via a monolayer of immobilized Azurin
(Az) was examined by conducting probe atomic force microscopy (CP-AFM), both as function of
temperature (248 - 373K) and of applied tip force (6-12 nN). By varying both temperature and force
in CP-AFM, we find that the ETp mechanism can alter with a change in the force applied via the tip
to the proteins. As the applied force increases, ETp via Az changes from temperature-independent to
thermally activated at high temperatures. This is in contrast to the Cu-depleted form of Az (apo-Az),
where increasing the applied force causes only small quantitative effects, that fit with a decrease in
electrode spacing. At low force ETp via holo-Az is temperature-independent and thermally activated
via apo-Az. This observation agrees with macroscopic-scale measurements, thus confirming that the
difference in ETp dependence on temperature between holo- and apo-Az is an inherent one that may
reflect a difference in rigidity between the two forms. An important implication of these results,
which depend on CP-AFM measurements over a significant temperature range, is that for ETp
measurements on floppy systems, such as proteins, the stress applied to the sample should be kept
constant or, at least controlled during measurement.
Keywords: Nanometer scale; conductivity; Azurin; biomolecular electronics; Arrhenius activation
energy; tunneling; electron transport
2
Azurin (Az) is an electron-mediating protein, functional in the bacterial energy conversion system,
which has been studied extensively, mostly by spectroscopic1-3 and electrochemical measurements4-
10 in solution, as a model system for electron transfer (ET) via proteins. Solid-state electron
transport (ETp) measurements have been reported on several types of proteins,11-13 including Az,
which has been investigated quite intensively, both by nano-scale techniques14-19 and by macro-scale
electrodes.20-23 Within the latter context we reported recently macroscopic scale ETp measurements
via Az monolayers,21 which showed temperature-independent currents via holo-Az. ETp via the Cu-
depleted form of Az, apo-Az, was found to be thermally-activated at T > 180K and temperature-
independent at T < 180K.
Investigating these current-voltage measurements with nano-scale contacts, may scrutinize two
important issues: (i) While the observed temperature independence of the currents through a 3.5 nm
thick protein (see below) was quite remarkable, we sought a different measurement method, less
likely to be influenced by possible defects (in the monolayer), to assure that this is indeed an
intrinsic property of Az. (ii) Use of conducting probe atomic force microscopy, CP-AFM, allows to
gain information on the ETp under different pressure applied to the protein. Up to now, only a
simple connection between tunneling currents and the force, applied to an organic monolayer was
found.24, 25 With proteins (and likely also for other soft materials), increasing the applied force may
cause structural changes, which should be considered, in addition to a simple compression that
increases current flow. By combining temperature dependent current transport measurements and the
unique AFM capability of force control on a nano-scale contact area, we can use ETp as an indicator
for force-dependent structural changes in proteins. These force-dependent measurements can also
provide comparison with macroscopic measurements, where only a negligible gravity force and a
constant adhesive force (both from the contact pad) act on the proteins. While, in principle, force
control in macroscopic measurement on a monolayer is possible, its use is hampered by the fact that
even slight variation in monolayer uniformity may cause penetration of the protein monolayer by the
electrode.
3
In a series of experiments, Davis and coworkers observed ETp through Az by CP-AFM.14, 16-19
From fits of measured I-V curves to Simmons’ non-resonant tunneling model they extracted values
for barrier height and length of the Az monolayer at different applied tip forces (at room
temperature). Along with ETp experiments via Az, current-voltage CP-AFM measurements of other
protein monolayers in a solid state environment have been reported, and the data were fitted to
different ETp models (for example Fowler-Nordheim tunneling26).15, 27, 28 CP-AFM allows
measuring ETp through a limited number of molecules, located between the tip and the surface, or
attached to the tip, to the surface or to both. The effective radius of the AFM tip limits the contact
area of the measurement and in the present work we estimate that we measure up to ~50 Az
molecules. Smaller contact area increases the probability of resolving small defect-free areas for
measurement. At the same time we sacrifice sensitivity (lower currents), compared to use of the
much more sensitive macroscopic electrode measurements that average over a large area. The two
approaches can, thus, be viewed as complementary ones.
According to Marcus theory,29 kET, the rate constant for electron transfer, ET between a donor (D)
and an acceptor (A), is given by:
(1)
where Ea is the activation energy of the process, T is the absolute temperature, kB is Boltzmann’s
constant and HD-A, is the degree of electronic coupling between D and A, which also depends on lD-A
the separation distance between them. While in ET studies lD-A is the commonly varied parameter,
and varying the temperature maybe problematic over a wide range (also in view of solvent
presence), for ETp the opposite situation holds. The Az is sandwiched between two electrodes and,
in our set-up is covalently attached to the bottom electrode by disulfide bridge through its cysteine
residue (Cys3 or Cys26). This configuration introduces a defined and fixed distance between the two
electrodes, l =3.5 nm, although in CP-AFM measurements one can change the tip-surface distance
by applying different probe forces.30 Using molecular dynamics simulations, Davis and coworkers
proposed that the secondary structure of the protein changes drastically upon applying increasing
4
2()exp(/)ETDADAaBkHlEkTforces to the protein layer.19 Accordingly, as will be shown and discussed below, at different probe
forces, we can expect the medium that separates the tip and the surface to be affected. This poses a
problem for certain nanoscopic measurements, such as those done by dynamic scanning tunneling
microscopy (STM),31 because the protein will be mechanically stressed and, possibly structurally
affected, during most of the measurement. Here we report temperature- and force-dependent CP-
AFM of holo- and apo-Az, as a monolayer, in a solid state-like junction over the temperature range
of 248-373K and applied forces of 6-12 nN.
RESULTS
In the employed configuration, a smooth (r.m.s. roughness of ~0.7 nm over 4×4 m scan) and
continuous Au layer is deposited on an H-terminated Si substrate. A self-assembled monolayer of
holo- or apo-Az was coupled covalently to the Au substrate via S-Au bonding between the Au
surface and the relatively exposed cysteine thiolate of Az (Fig. 1). To confirm the establishment of
the protein monolayer, topographic (amplitude) and phase images were measured in semi-contact
mode with different amplitude reduction or amplitude set point ratio (97%, 94% and 91%,
respectively) (Fig. 2). A lower set point ratio corresponds to a higher tip force on the surface. The
relative influence of attractive (adhesive) forces changes as the set point ratio is lowered, particularly
on compliant surfaces, leading to modification of topography and phase images.32-34 As we lowered
the set point ratio, i.e., increased the tip force, we clearly observed a change in the topography
images, and corresponding changes in the phase images. The observed changes in the topography
and phase images with different tip forces suggest the presence of a soft layer, i.e. proteins. To
assess the thickness of the protein monolayer, a square area was scanned in the contact mode with a
large feedback force (160 nN). The applied force is sufficiently large to scratch away the monolayer.
Following the contact mode scratching procedure we reverted to semi-contact mode to re-scan over
a larger area, centered around the resulting trough (Supporting Information, Figure S1 a). The
obtained image demonstrates that the probe removed the protein over a square area. The cross-
5
section (Supporting Information, Figure S1 b) shows that the thickness of the protein layer is about
3.5 nm, the length of Az established by its three-dimensional structure determination.
The measured I-V of holo- and apo-Az as a function of applied force (conducted at room
temperature), is shown in Figure 3. The currents obtained for both proteins increased with probe
force. For the holo-Az junctions at 0.5 V bias, the currents ranged between 0.1-1 nA for tip forces of
6 – 15 nN, in agreement with the values reported by Zhao and Davis.14, 19 Clear shorts were observed
everywhere with > 40 nN tip force, suggesting the tip pinches through the protein layer at this force.
The currents across apo-Az junctions spanned a range of 0.002 – 2 nA for the same force variation.
The higher sensitivity of apo-Az to the applied tip force suggests that it is more flexible than holo-
Az. Indeed, the crystal structure of apo-Az, though very similar to that of holo-Az, indicates more
flexibility,35-37 particularly at the, for apo-Az empty, metal binding site. If a low force is applied at
room temperature, the observed current across apo-Az junctions at 1 V was an order of magnitude
lower than that across the holo-Az junctions (Fig 3, black curve). This difference resembles that
obtained in our macroscopic measurements21 and is qualitatively consistent with results of
conduction measurements via another metal-containing protein (Ferritin) by CP-AFM27, STM38 or,
for Az in a three-terminal configuration.39 This decrease in current magnitude highlights the role of
the Cu ion as an efficient ET and ETp mediator in Az.
Figure 4 shows the current-voltage characteristics of holo-Az (Figs. 4a,b) and the natural
logarithm of the currents for voltage sweeps from -1 V to +1 V for apo-Az junctions (Fig. 4c,d) as a
function of temperature. The I-V curves were measured over more than 30 random spots for each
temperature at constant probe force (6 nN). For each temperature we took the average of all I-V
curves, excluding those exhibiting shorting or insulating behavior (~20%) The distribution of current
intensities was normal (inset of Fig. 4a); average currents were used for further analysis. The I-V
measurements were performed at each temperature only after the system had remained stable for at
least 30 minutes. As can be seen in figure 4, holo-Az and apo-Az exhibit distinct ETp characteristics
as a function of temperature. The I-V curves of the holo-Az junctions remained essentially invariant
with temperature from 248 to ~370 (Fig. 4a). The measured current via the junctions at +0.5 V and –
6
0.5 V as a function of the inverse temperature (Fig. 4b) clearly demonstrates a temperature-
independent ETp process across holo-Az. Thus, the CP-AFM results, obtained with a relatively low
tip force of 6 nN, confirm our macroscopic ETp observations.21 In the CP-AFM measurements, we
observed a constant adhesion force of 5 nN, which was measured as the snap-out force during
retraction of the AFM tip. Thus, the total load applied on the proteins is 11 nN. In the macroscopic
measurement using an Au pad, the force, exerted by the Au on the proteins, is mainly governed by
the intrinsic adhesion force. Symmetry considerations imply that this force is similar to that of this
intrinsic metal tip/protein adhesion. Since in the macroscopic scale measurements there is no
significant additional external load, the effective pressure applied to each protein molecule is less
than half that in the nano-scale measurements.
The role of the Cu redox center in the ETp process across the protein was measured on a
monolayer of apo-Az. Similar (optical) heights were deduced from ellipsometry for apo-Az and
holo-Az (18Å), comparable to the observed height on Si surfaces20, corresponding to similar packing
of the monolayer on the Au surface. The junctions of apo-Az were prepared, similarly to those of
holo-Az, and the same AFM tips were used to measure the I-V characteristics of the two proteins.
Figure 4c shows the I-V curves of the apo-Az junction on a logarithmic current scale at different
temperatures, in the range of 248-368K. Figure 4d shows the logarithm of the currents through the
junctions at +0.5 V as a function of inverse temperature. A linear dependence is clearly observed.
Thus, removing the Cu ion changes the dominant mechanism of ETp across the protein from
temperature-independent to the more common thermally activated type , in agreement with the
macroscopic ETp results.21
Upon raising the temperature further we observed a sharp irreversible decrease of the currents for
apo- and holo-Az junctions (marked in circle in Fig. 4b and d), this is in contrary to the reversible
temperature dependent that we observe (by heating and cooling) in the monolayer junction at lower
temperatures. The most likely reason for the irreversible decrease in the currents is the denaturation
of the protein. The reported Tm values of apo- and holo-Az in solution are 335K and 355K
respectively,40 which are slightly lower than the observed denaturation in the macro- and nano-scale
7
measurement. Higher stability, consequently leading to higher Tm has also been observed previously
in other dry proteins.41 In our macroscopic scale experiments these current drops appear at 350K and
360K for apo and holo-Az, respectively.21 The discrepancies of 5-15K may be due to the different
coupling modes to the electrodes, which may affect the stability of the protein as well.
The effect of applied tip force on the current as a function of temperature is shown in Figures 5a
and b, which give the currents at +0.5 V applied bias on the tip as a function of inverse temperatures
for holo- and apo-Az, respectively, at 6, 9 and 12 nN tip force. As shown in Fig. 3 for room
temperature currents, increasing the force, leads to higher current. However, we find a distinct
difference in the temperature dependence of the current between apo- and holo-Az, at different tip
forces. While currents via apo-Az increase with increasing tip force, as expected from a decrease in
the tip-substrate distance, their temperature dependence remains similar. By fitting the curves in Fig.
5b to the Arrhenius equation, we can estimate the ETp activation energy, Ea, of apo-Az to be
600±100 meV at all of the applied forces. In contrast, the currents via holo-Az change from
temperature-independent at 6 nN to thermally activated at higher applied forces (9 and 12 nN) at
temperatures > 310K. Below 310K the currents through holo-Az remain temperature-independent at
all of the applied forces, but currents are higher for higher applied force, which can be the result of a
decrease in tip-substrate distance. These differences between apo- and holo-Az can also be seen in a
plot of the measured currents as a function of the applied tip force (Fig. 6), obtained by combining
simultaneous force and current traces as a function of time by high speed data acquisition in Peak
Force TUNATM mode. Our results show that while the slopes of the currents as a function of the tip
force for apo-Az (Fig. 6a) remain similar at different temperatures, the slope for holo-Az increased
significantly with temperature.
The above mentioned irreversible drop in current upon heating was observed at similar
temperatures for both holo-Az and apo-Az, with 6, 9 and 12 nN tip force applied. This similarity
suggests that any structure distortion of the protein caused by a tip force of up to 12 nN is minor,
compared to the change involved in denaturation.
8
DISCUSSION
Comparison of current densities observed in macroscopic and nanoscopic contact measurements
Thermally activated ETp was previously observed in CP-AFM studies via > 4 nm long conjugated
molecules (with temperature-independent behavior for shorter molecules).42, 43 The Arrhenius plots
of the currents as a function of inverse temperature observed in those experiments were interpreted
in terms of a hopping ETp mechanism. This may well be the dominant ETp mechanism via apo-Az.
While Az, with or without its Cu redox centre, is not a conjugated system, apo-Az behaves
qualitatively in a similar fashion to conjugated molecules (“molecular wires”), but with measured
currents of 10’s of pA, rather than the A currents that flow through conjugated molecules of similar
molecular length at 0.5 V applied bias.43 The contact area for these conjugated molecules was
estimated at 50 nm2.43 We can estimate our contact area by using the following relation derived from
Herzian contact mechanics:24
where r is the tip radius, Peff is the load and E* is the effective modulus, which we estimate to be
~100 MPa.44, 45 From this calculation we assess the contact area in our studies to be ~450 nm2. After
normalization of the contact area differences, we find the current densities for both holo- and apo-Az
to be 5-6 orders of magnitude lower than those of conjugated molecules. In comparison, an
insulating layer of alkyl chains of similar thickness, at comparable force and tip radius should pass
10-18A (extrapolating from CP-AFM measurement on CH3(CH2)11SH, ~500 pA at 0.5 V, and using
=1.1 A-1 25). This observation by itself makes proteins, in terms of conductivity, more akin to
molecular wires than to insulators.
In general, the CP-AFM current-temperature behavior of Az with nano-scale electrode contact is
consistent with our previously reported macroscopic results,21 in that they both show temperature-
independent ETp of holo-Az and thermally activated ETp of apo-Az. There are however some
differences between the CP-AFM and the macroscopic conductance measurements25 in the
9
normalized current densities across the proteins and in the activation energy, Ea, of apo-Az that was
derived from the data.
Current Densities
From the nano-scale measurements (AFM tip radius of 20 nm) we calculate for holo-Az junctions
a current density of ~100 A/cm2 at 1 V, while in the macroscopic measurements (contact area of 0.2
mm2) it was ~3x10-3 A/cm2, namely five orders of magnitude lower. We ascribe this difference to
differences in the way the proteins are immobilized between the two electrodes in the two
measurements. In the macroscopic conductance measurements the substrate was a p++-Si surface
with a 1 nm thick layer of silicon oxide, with, as linker, a self-assembled monolayer of a C3 organo-
silane on top of it. The linker molecules were covalently bound to the Az proteins via a disulphide
bond.20 The top electrode was a macroscopic pad of gold (~0.2 mm2), deposited by the lift-off, float-
on (LOFO) technique,46 to contact the protein layer. Thus, in the macroscopic measurements the two
contacts were separated not only by the protein monolayer, but also by two additional insulating
layers, silicon oxide and organic linker, with a combined thickness of ~16 Å). These two additional
insulating layers will decrease the currents by some five orders of magnitude, assuming a current
decay factor (for insulating molecular layers) of = 0.7 A-1.12
Thermal activation energy for ETp via apo-Az
The distinct experimental setups may also be the cause for the different calculated ETp activation
energies of apo-Az, i.e., 320 meV from the macroscopic measurements30 and ~600 meV from the
CP-AFM measurements. Because the holo-Az results exhibited the same temperature-independent
behavior in the two different experimental setups, it is very unlikely that the different electrodes are
the cause for this difference in activation energy. A possible reason is that the covalent S-S bond
between one of the two exposed cysteines and SH group of the linker in the macroscopic
measurements is stronger, than the Au-S bond that binds the protein to the Au substrate in the CP-
AFM measurements. The smaller coupling to the electrode might allow additional vibration modes
10
at the Au/protein interface, which are hindered in the case of the disulfide bridge bond that holds the
protein in the macroscopic configuration.
Tip pressure dependence of the ETp
A striking result of this study is the observed distinct temperature dependence of the current
through holo-Az as a function of applied force (Figs. 5 and 6). As discussed previously, the higher
sensitivity to the applied tip force of currents through apo-Az (at room temperature) than through
holo-Az, (Fig. 3) is interpreted as being a result of the increased flexibility of apo-Az, compared to
holo-Az. This increased flexibility of apo-Az causes the tip-surface separation distance to be smaller
at high forces than in the holo-Az surface, which explains the larger increase in currents with
increasing force via apo-Az. In other words, Young’s modulus of apo-Az protein is lower than holo-
Az, leading to a more significant change in the separation distance between the tip and the gold
substrate, due to protein compression, hence leading to more pronounced changes in currents
between the electrodes.
An increase in the applied tip force on top of the holo-Az monolayer results in the AFM tip
pushing toward the Cu binding site (which is located on the tip-side of the junction as shown in the
scheme of Fig. 1). Marshall et al.47 have shown that minor alterations of the hydrogen-bonding
network in the second coordination sphere of the Cu site may cause pronounced changes in the ET
properties of holo-Az, which consequently can change its flexibility.48-50 The change (at T > 310 K)
from temperature-independent to thermally-activated ETp via holo-Az suggests a relation between
the protein structure and flexibility, and the mechanism of ETp through it.51 The suggested
correlation between flexibility and conductivity47 has been indicated by measurements of ETp via
peptide nucleic acid monolayers, where attenuation in currents was observed for structures
comprising the same sequence, with and without methylation, a modification that increases the
structural rigidity.52 Therefore, we suggest that the protein’s flexibility (in the case of holo-Az),
together with the temperature of the system (for apo-Az) are dominant parameters that control both
the mechanism and the efficiency of ETp via the protein.
11
The impact of applied forces on the ETp characteristics of proteins suggests that the compressive
and tensile stress applied to the examined protein need to be taken into consideration when its
electrical conduction properties are investigated. , In similar ETp measurements that were conducted
via alkyl chains, currents began to increase only at > 12nN tip force,25 suggesting a more rigid
structure than that of proteins. Furthermore, the force applied by the tip on an alkyl chain has
apparently a simpler effect, than that on proteins, where more complex secondary structures can be
affected. Thus, consideration of the applied force appears to be particularly relevant for nanoscopic
approaches, such as STM and CP-AFM, for measurements of proteins, as shown here.
CONCLUSIONS
CP-AFM measurements of protein monolayers show that, at low tip-force, ETp through holo-Az is
temperature-independent over a significant range of temperatures (248-373K), while for apo-Az the
process is thermally activated. This observed difference is in line with our earlier macroscopic ETp
results. As the tip-force increases, ETp through holo-Az changes from temperature-independent to
thermally activated at higher temperatures (> 310K). It is likely that the mechanism of ETp through
apo-Az does not change with pressure, as the thermal activation energy does not change. The
currents, though, increase with increasing tip force, an effect that can by ascribed to decreases of
through-space ETp gaps (by compression of tunneling pathways) in the protein, consistent with
results of Meier et al.53 The results obtained with holo-Az do suggest a pressure-induced change in
ETp mechanism at temperatures >310K, which may be due to a pressure-induced change in protein
structure. Even though this change is likely to be minor, because it does not significantly alter the
denaturation temperature, it does manifest itself in ETp. Thus, this result illustrates the remarkable
sensitivity of ETp to both relatively minor structural changes, as well as to major changes in the
protein (denaturation). The strong relation between protein flexibility and applied force, and its
impact on the temperature-dependence of ETp via Az, indicates the importance of the applied forces
for current measurements in scanning microscopy configurations (cf., review on biomolecule
conductance11 and report on force-dependent conductivity14).
12
Finally, our observation directly points to a general issue that is worth considering: force-mediated
effects in nano-scale electrical measurements. The applied forces should be the lowest that still
assure reproducible ETp. Because a change in applied force such as the one demonstrated here, can
affect ETp, changes in force during ETp measurements, should be avoided or their effect should be
considered. This consideration is especially important for ETp measurements through proteins
because of the relative ease by which they can change conformation, compared to more rigid
molecules, such as, e.g., conjugated or alkyl ones. The combination of controlled temperature and
environment, together with a defined (and reported) tip diameter and force are critical parameters
that define the initial conditions of the system, and, hence, the output of the measurement.
13
METHODS
Preparation of the substrates. Silicon wafers (p type, boron doped, <100> single side polished, <
0.001 ·cm), were sonicated for 2 min with ethyl acetate, acetone and ethanol. Immediately after
that, the wafer was etched for 1 min with 2% HF, washed with Milli-Q (18M) water and cleaned
again with fresh piranha solution for 20 min (7/3 v/v of H2SO4/H2O2) at 80°. After cleaning with
piranha the wafers were rinsed thoroughly with Mili-Q and etched with 2%HF again, resulting in a
Si-H surface termination. The wafers were immediately stored in a container, filled with nitrogen.
The samples were loaded into an e-beam evaporator and, when a vacuum of < 5*10-6 mbar was
reached, 2 nm of Cr (serving as an adhesion layer) followed by 50 nm of Au were evaporated with a
deposition rate of 1Å/s. The gold-coated Si samples were then cut into 1*1 cm2 slides, cleaned for 10
min with UV/ozone treatment followed by a 30 min immersion in ethanol.54
Preparation and characterization of the proteins. Azurin was isolated from Alcaligenes faecalis
by the method of Ambler and Wynn.55 Apo-Az solution was prepared as described.20 Holo-Az and
apo-Az monolayers were prepared by immersing the gold substrates in a ~1 mg/mL solution of
azurin in 50 mM ammonium acetate (NH4Ac) buffer (pH 4.6) for 2 h followed by rinsing in clean
H2O and drying under a fine nitrogen stream. The protein monolayers were characterized by
ellipsometry that yielded 1.8 nm optical thickness, which corresponds to values obtained in our
previous studies of Az on Si surfaces.20 The actual thickness is 3.5 nm, as found by AFM (Supp.
Info., Fig. S1b)
AFM imaging. The topography of the self-assembled monolayer of proteins was characterized by
AFM in semi contact mode under N2 purge. A Solver P47 SPM system (ND-MDT, Zelenograd
Russia) and Pt coated Si probes (NSC36, 75kHz, 0.6 N/m, MIKROMASCH) were used. The
topography images and the phase images were taken simultaneously at a scan rate of 1 Hz. The
applied force from the tip on the proteins, during the topography imaging, is at least one order of
magnitude smaller than those used for CP-AFM measurements. (supporting information)
14
CP-AFM measurements as a function of temperature. The CP-AFM measurements as a function
of temperature were performed with a Multimode/Nanoscope V system (Bruker-Nano, Santa
Barbara, CA USA) under constant N2 flow purge. All-metal Pt AFM probes with nominal force
constant of 0.8 N/m (25PT300B, Rocky Mountain Nanotechnology, Salt Lake City Utah USA) and
tip radius of 20 nm, were used for the CP-AFM measurements. The probes were brought into contact
with the self-assembled protein monolayer using a constant force feedback (contact mode). Given
the tip radius, ~50 proteins were measured and averaged, so our temperature and force dependent
measurements are less affected by the drifting and penetration of the proteins than in the case of a
single protein. A tip force of 6 nN was used to avoid changes in the protein, following ref.19
Temperature was controlled by a Bruker heater/cooling system that uses an inert atmosphere flow,
rather than a closed vacuum system as in the case of our macroscopic measurements. With this set -
up we can obtain current voltage (I-V) curves between 248K and 373K. At temperatures below 273K
(0º C) the protein surface was monitored constantly to verify that no ice is formed, ascertaining the
low moisture level in the system. For each temperature, the I-V measurements were performed only
after the system had stabilized and remained stable for at least 30 minutes. 30 I-V curves over an
area of 1×1 µm2 were taken and averaged for each temperature. The drifting of the tip position will
not affect our measurements. Raw data of I-V curves from proteins samples at room temperature
were plotted in Figures S2 (Supporting Information) to show the variation of the I-V curves.
Whenever drastic current drops occurred, the tip was cleaned, using a high voltage pulse (-10 to+10
V in 0.1 sec). The tip force was constant during the measurements.
AFM Peak Force TUNA Mode. An extension module for current measurement was used to enable
the current mapping under AFM Peak Force (PF-TUNATM) mode. In Peak Force TUNA mode, the
probe was cycled in and out of contact with the surface at 1 or 2 kHz, while the tip is scanned across
the sample at a rate of 1 Hz per scan line. The fast data acquisition, coupled with feedback loop
control, the maximum force on the tip for each individual cycle. Current, force and other mechanical
properties are recorded during these controlled tip-sample contact cycles. Peak current is recorded at
15
maximum force applied to the sample (Peak Force set point). Measurements were performed in both
imaging mode and spectroscopy mode. In the imaging mode, a current map is acquired
simultaneously with topography at a tip bias of +0.5 V with a tip Peak Force of 6, 9 nN and 12 nN.
The currents from the image were averaged for each such mapping at a given temperature. In
spectroscopy mode, the tip was ramped into and out of the surface at high cycling frequency (1 kHz),
while recording simultaneously both force and current as function of time. The current force curve
was plotted by combing the two simultaneous traces.
Acknowledgments WL thanks Daniel Frisbie and Liang Luo (Un. Minnesota) for helpful
guidance with, and discussions on CP-AFM measurement procedures. LS thanks the Eshkol
program for financial support. NA thanks the Clore program for financial support. We thank the
Minerva Foundation (Munich), the Kimmelman center for Biomolecular Structure and Assembly,
the Kimmel centre for Nanoscale Science and the Grand Centre for Sensors and Security for partial
support. MS holds the Katzir-Makineni chair in Chemistry. DC holds the Schaefer Chair in Energy
Research.
Supporting Information Available: additional information about AFM imaging, AFM tip
scratching procedure, -protein thickness measurements and raw data of I-V curves for Az and apo-
Az at room temperature. This material is available free of charge via the Internet at
http://pubs.acs.org.
16
References
1.
Farver, O.; Pecht, I., Elucidation of Electron-Transfer Pathways in Copper and Iron Proteins
by Pulse Radiolysis Experiments. In Progress in Inorganic Chemistry, Vol 55, Karlin, K. D., Ed.
2007; Vol. 55, pp 1-78.
Gray, H. B.; Winkler, J. R., Electron tunneling through proteins. Q. Rev. Biophys. 2003, 36
2.
(3), 341-372.
3.
Regan, J. J.; Dibilio, A. J.; Langen, R.; Skov, L. K.; Winkler, J. R.; Gray, H. B.; Onuchic, J.
N., Electron-tunneling in azurin - the coupling across a beta-sheet. Chemistry & Biology 1995, 2 (7),
489-496.
4.
Chi, Q. J.; Zhang, J. D.; Andersen, J. E. T.; Ulstrup, J., Ordered assembly and controlled
electron transfer of the blue copper protein azurin at gold (111) single-crystal substrates. J. Phys.
Chem. B 2001, 105 (20), 4669-4679.
5.
Chi, Q. J.; Zhang, J. D.; Nielsen, J. U.; Friis, E. P.; Chorkendorff, I.; Canters, G. W.;
Andersen, J. E. T.; Ulstrup, J., Molecular monolayers and interfacial electron transfer of
Pseudomonas aeruginosa azurin on Au(111). J. Am. Chem. Soc. 2000, 122 (17), 4047-4055.
6.
Gaigalas, A. K.; Niaura, G., Measurement of electron transfer rates between adsorbed azurin
and a gold electrode modified with a hexanethiol layer. J. Colloid Interface Sci. 1997, 193 (1), 60-
70.
7.
Jeuken, L. J. C.; Armstrong, F. A., Electrochemical origin of hysteresis in the electron-
transfer reactions of adsorbed proteins: Contrasting behavior of the "blue" copper protein, azur in,
adsorbed on pyrolytic graphite and modified gold electrodes. J. Phys. Chem. B 2001, 105 (22), 5271-
5282.
8.
Alessandrini, A.; Corni, S.; Facci, P., Unravelling single metalloprotein electron transfer by
scanning probe techniques. Phys. Chem. Chem. Phys. 2006, 8 (38), 4383-4397.
9.
Friis, E. P.; Andersen, J. E. T.; Madsen, L. L.; Moller, P.; Ulstrup, J., In situ STM and AFM
of the copper protein Pseudomonas aeruginosa azurin. J. Electroanal. Chem. 1997, 431 (1), 35-38.
10.
Andolfi, L.; Bruce, D.; Cannistraro, S.; Canters, G. W.; Davis, J. J.; Hill, H. A. O.; Crozier,
J.; Verbeet, M. P.; Wrathmell, C. L.; Astier, Y., The electrochemical characteristics of blue copper
protein monolayers on gold. J. Electroanal. Chem. 2004, 565 (1), 21-28.
11.
Shinwari, M. W.; Deen, M. J.; Starikov, E. B.; Cuniberti, G., Electrical Conductance in
Biological Molecules. Adv. Funct. Mater. 2010, 20 (12), 1865-1883.
Ron, I.; Pecht, I.; Sheves, M.; Cahen, D., Proteins as Solid-State Electronic Conductors. Acc.
12.
Chem. Res. 2010, 43 (7), 945-953.
13.
Das, R.; Kiley, P. J.; Segal, M.; Norville, J.; Yu, A. A.; Wang, L. Y.; Trammell, S. A.;
Reddick, L. E.; Kumar, R.; Stellacci, F.; Lebedev, N.; Schnur, J.; Bruce, B. D.; Zhang, S. G.; Baldo,
M., Integration of photosynthetic protein molecular complexes in solid-state electronic devices.
Nano Lett. 2004, 4 (6), 1079-1083.
14.
Zhao, J. W.; Davis, J. J., Force dependent metalloprotein conductance by conducting atomic
force microscopy. Nanotechnology 2003, 14 (9), 1023-1028.
15.
Davis, J. J.; Peters, B.; Xi, W., Force modulation and electrochemical gating of conductance
in a cytochrome. J. Phys.-Condes. Matter 2008, 20 (37).
16.
Davis, J. J.; Wang, N.; Morgan, A.; Zhang, T. T.; Zhao, J. W., Metalloprotein tunnel
junctions: compressional modulation of barrier height and transport mechanism. Faraday Discuss.
2006, 131, 167-179.
17.
Davis, J. J.; Wrathmell, C. L.; Zhao, J.; Fletcher, J., The tunnelling conductance of
molecularly ordered metalloprotein arrays. J. Mol. Recognit. 2004, 17 (3), 167-173.
18.
Zhao, J. W.; Davis, J. J., Molecular electron transfer of protein junctions characterised by
conducting atomic force microscopy. Colloids Surf., B 2005, 40 (3-4), 189-194.
19.
Zhao, J. W.; Davis, J. J.; Sansom, M. S. P.; Hung, A., Exploring the electronic and
mechanical properties of protein using conducting atomic force microscopy. J. Am. Chem. Soc.
2004, 126 (17), 5601-5609.
17
Good, R. H. J.; Muller, E. W., Handbuch der Physik. Springer Verlag: Berlin, 1956; Vol.
20.
Ron, I.; Sepunaru, L.; Itzhakov, S.; Belenkova, T.; Friedman, N.; Pecht, I.; Sheves, M.;
Cahen, D., Proteins as Electronic Materials: Electron Transport through Solid-State Protein
Monolayer Junctions. J. Am. Chem. Soc. 2010, 132 (12), 4131-4140.
21.
Sepunaru, L.; Pecht, I.; Sheves, M.; Cahen, D., Solid-State Electron Transport across Azurin:
From a Temperature-Independent to a Temperature-Activated Mechanism. J. Am. Chem. Soc. 2011,
133 (8), 2421-2423.
22.
Bizzarri, A. R.; Andolfi, L.; Taranta, M.; Cannistraro, S., Optical and electronic coupling of
the redox copper Azurin on ITO-coated quartz substrate. Biosensors & Bioelectronics 2008, 24 (2),
204-209.
23. Mentovich, E. D.; Belgorodsky, B.; Richter, S., Resolving the Mystery of the Elusive Peak:
Negative Differential Resistance in Redox Proteins. J. Phys. Chem. Lett. 2011, 2 (10), 1125-1128.
24.
Engelkes, V. B.; Beebe, J. M.; Frisbie, C. D., Analysis of the causes of variance in resistance
measurements on metal-molecule-metal junctions formed by conducting-probe atomic force
microscopy. J. Phys. Chem. B 2005, 109 (35), 16801-16810.
25. Wold, D. J.; Frisbie, C. D., Fabrication and Characterization of Metal-Molecule-Metal
Junctions by Conducting Probe Atomic Force Microscopy. J. Am. Chem. Soc. 2001, 123 (23), 5549-
5556.
26.
XXI.
Xu, D.; Watt, G. D.; Harb, J. N.; Davis, R. C., Electrical Conductivity of Ferritin Proteins by
27.
Conductive AFM. Nano Lett. 2005, 5 (4), 571-577.
28.
Stamouli, A.; Frenken, J. W. M.; Oosterkamp, T. H.; Cogdell, R. J.; Aartsma, T. J., The
electron conduction of photosynthetic protein complexes embedded in a membrane. FEBS Lett.
2004, 560 (1), 109-114.
29. Marcus, R. A.; Sutin, N., Electron transfers in chemistry and biology. Biochim. Biophys. Acta
1985, 811 (3), 265-322.
30.
In our macroscopic measurements the force, applied by the contact pad, which is primarily
due to adhesion forces, translates into a fixed force applied to the protein.
31.
Tao, N. J., Probing potential-tuned resonant tunneling through redox molecules with
scanning tunneling microscopy. Physical Review Letters 1996, 76 (21), 4066-4069.
32.
Bar, G.; Thomann, Y.; Brandsch, R.; Cantow, H. J.; Whangbo, M. H., Factors affecting the
height and phase images in tapping mode atomic force microscopy. Study of phase -separated
polymer blends of poly(ethene-co-styrene) and poly(2,6-dimethyl-1,4-phenylene oxide). Langmuir
1997, 13 (14), 3807-3812.
33.
Kuhle, A.; Sorensen, A. H.; Bohr, J., Role of attractive forces in tapping tip force
microscopy. Journal of Applied Physics 1997, 81 (10), 6562-6569.
34.
San Paulo, A.; Garcia, R., Amplitude, deformation and phase shift in amplitude modulation
atomic force microscopy: a numerical study for compliant materials. Surface Science 2001, 471 (1-
3), 71-79.
35.
Shepard, W. E. B.; Kingston, R. L.; Anderson, B. F.; Baker, E. N., Structure Of Apo-Azurin
From Alcaligenes-Denitrificans At 1.8-Angstrom Resolution. Acta Crystallogr., Sect. D: Biol.
Crystallogr. 1993, 49, 331-343.
36.
Nar, H.; Messerschmidt, A.; Huber, R.; Vandekamp, M.; Canters, G. W., Crystal-Structure
Of Pseudomonas-Aeruginosa Apo-Azurin At 1.85 Angstrom Resolution. FEBS Lett. 1992, 306 (2-
3), 119-124.
37.
Vandekamp, M.; Canters, G. W.; Wijmenga, S. S.; Lommen, A.; Hilbers, C. W.; Nar, H.;
Messerschmidt, A.; Huber, R., Complete Sequential H-1 And N-15 Nuclear-Magnetic-Resonance
Assignments And Solution Secondary Structure Of The Blue Copper Protein Azurin From
Pseudomonas-Aeruginosa. Biochemistry 1992, 31 (42), 10194-10207.
38.
Rakshit, T.; Banerjee, S.; Mukhopadhyay, R., Near-Metallic Behavior of Warm Holoferritin
Molecules on a Gold(111) Surface. Langmuir 2010, 26 (20), 16005-16012.
18
39. Maruccio, G.; Marzo, P.; Krahne, R.; Passaseo, A.; Cingolani, R.; Rinaldi, R., Protein
conduction and negative differential resistance in large-scale nanojunction arrays. Small 2007, 3 (7),
1184-1188.
40.
Engeseth, H. R.; McMillin, D. R., Studies of thermally induced denaturation of azurin and
azurin derivatives by differential scanning calorimetry: evidence for copper selectivity. Biochemistry
1986, 25 (9), 2448-2455.
41.
Shen, Y.; Safinya, C. R.; Liang, K. S.; Ruppert, A. F.; Rothschild, K. J., Stabilizat ion of the
membrane protein bacteriorhodopsin to 1400C in two-dimensional films. Nature 1993, 366 (6450),
48-50.
42.
Luo, L.; Choi, S. H.; Frisbie, C. D., Probing Hopping Conduction in Conjugated Molecular
Wires Connected to Metal Electrodes. Chem. Mater. 2010, 23 (3), 631-645.
43.
Choi, S. H.; Frisbie, C. D., Enhanced Hopping Conductivity in Low Band Gap Donor -
Acceptor Molecular Wires Up to 20 nm in Length. J. Am. Chem. Soc. 2010, 132 (45), 16191-16201.
Atsushi, I., Local rigidity of a protein molecule. Biophys. Chem. 2005, 116 (3), 187-191.
44.
45. WolframAlpha.
http://www.wolframalpha.com/input/?i=biological+material&lk=1&a=ClashPrefs_*MaterialClass.B
iologicalMaterial-. (accessed July 17, 2012).
46.
Vilan, A.; Cahen, D., Soft contact deposition onto molecularly modified GaAs. Thin metal
film flotation: Principles and electrical effects. Adv. Funct. Mater. 2002, 12 (11-12), 795-807.
47. Marshall, N. M.; Garner, D. K.; Wilson, T. D.; Gao, Y. G.; Robinson, H.; Nilges, M. J.; Lu,
Y., Rationally tuning the reduction potential of a single cupredoxin beyond the natural range. Nature
2009, 462 (7269), 113-116.
Benning, M. M.; Meyer, T. E.; Rayment, I.; Holden, H. M., Molecular Structure of the
48.
Oxidized High-Potential
Iron-Sulfur Protein
Isolated
from Ectothiorhodospira vacuolata.
Biochemistry 1994, 33 (9), 2476-2483.
49.
Palfey, B. A.; Basu, R.; Frederick, K. K.; Entsch, B.; Ballou, D. P., Role of Protein
Flexibility in the Catalytic Cycle of p-Hydroxybenzoate Hydroxylase Elucidated by the Pro293Ser
Mutant. Biochemistry 2002, 41 (26), 8438-8446.
50.
Battistuzzi, G.; Borsari, M.; Cowan, J. A.; Ranieri, A.; Sola, M., Control of Cytochrome c
Redox Potential: Axial Ligation and Protein Environment Effects. J. Am. Chem. Soc. 2002, 124 (19),
5315-5324.
51.
Beratan, D. N.; Skourtis, S. S.; Balabin, I. A.; Balaeff, A.; Keinan, S.; Venkatramani, R.;
Xiao, D. Q., Steering Electrons on Moving Pathways. Acc. Chem. Res. 2009, 42 (10), 1669-1678.
52. Wierzbinski, E.; de Leon, A.; Davis, K. L.; Bezer, S.; Wolak, M. A.; Kofke, M. J.; Schlaf,
R.; Achim, C.; Waldeck, D. H., Charge Transfer through Modified Peptide Nucleic Acids. Langmuir
2012, 28 (4), 1971-1981.
53. Meier, M.; van Eldik, R.; Chang, I. J.; Mines, G. A.; Wuttke, D. S.; Winkler, J. R.; Gray, H.
B., Pressure Effects on the Rates of Intramolecular Electron Transfer in Ruthenium -Modified
Cytochrome c. Role of the Intervening Medium in Tuning Distant Fe2+:Ru3+ Electronic Couplings.
J. Am. Chem. Soc. 1994, 116 (4), 1577-1578.
54.
Ron, H.; Matlis, S.; Rubinstein, I., Self-assembled monolayers on oxidized metals. 2. Gold
surface oxidative pretreatment, monolayer properties, and depression formation. Langmuir 1998, 14
(5), 1116-1121.
55.
Ambler, R. P.; Wynn, M., Amino-acid sequences of cytochromes C-551 from 3 species of
Pseudomonas. Biochem. J 1973, 131 (3), 485-498.
19
Figure Captions
Figure 1. Schematic (not to scale) of junction configuration employed in the ETp measurement via
the proteins. (PDB ID: 1AZU)
Figure 2. AFM topography (left) and phase (right) images of holo-Az with different amplitude set
point ratios (97%, 94% and 91%, respectively), corresponding to different forces applied on the
proteins.
Figure 3. Representative force-dependent I-V curves of (a) holo-Az and (b) apo-Az at room
temperature. The current increases with force from 6 nN to 15 nN at 1 V are 9.7 ± 2 times and 5.5
10-2 ± 3.102 times for holo-Az and apo-Az, respectively. See Suppl. Info. for a raw data set.
Figure 4. (a) Averaged I-V curves of holo-Az at temperatures from 248 to 373 K. The inset shows a
normal distribution of the currents taken at +1 V. (b) Currents of holo-Az at +0.5 V (black squares)
and –0.5 V (red triangles) as a function of inverse temperature. (c) Semi-logarithmic averaged I-V
curves of apo-Az at temperatures from 268 to 368 K. (d) Currents of apo-Az at +0.5 V as a function
of inverse temperature. The black curves in (a) and (c) are for the samples after denaturation. They
correspond to the circled points in (b) and (d), respectively. Error bars were calculated by the
standard deviation of I-V curve measurements for each temperature.
Figure 5. ETp temperature dependence for (a) holo-Az and (b) apo-Az junctions, shown as plots of
current at 0.5 V (logarithmic scale) vs. inverse temperature at 6, 9 and 12 nN applied tip force. Error
bars are based on the standard deviation of the “Peak Force” currents from current mapping.
Figure 6. Representative single measurement, current-force curves (logarithmic scale) at 0.5 V at
different temperatures for (a) apo-Az and (b) holo-Az junctions. The slopes of the logarithmic
current vs. force plots are 0.6 ± 0.1 for apo-Az for all temperatures, and 0.4 ± 0.1 at 288K and 0.9 ±
0.2 at 338 K for holo-Az. The currents at 288 K at the lowest forces reflect the sensitivity limit for
single measurements (10 pA).
20
Figure 1
Figure 2
97%
94%
91
21
(a)
(b)
Figure 3
(b)
(a)
(c)
358K
(d)
268
K
Figure 4
22
2.62.83.03.23.43.63.8100101102103104T (K) I (pA)1000/T (1/K)380360340320300280-1.0-0.50.00.51.0-0.4-0.20.00.20.40.6 I (nA)V (V)2.62.83.03.23.43.63.84.0-0.4-0.3-0.2-0.10.00.10.20.30.4 + 0.5 V - 0.5 V I (nA)1000/T (1/K)380360340320300280260240 -1.0-0.50.00.51.010-1100101102103104 I (pA)V (V)0.00.20.40.60.81.002468101214CountsI (nA)-1.0-0.50.00.51.0-1.0-0.50.00.51.0 6 nN 9 nN 12 nN 15 nNI (nA)V (V)-1.0-0.50.00.51.0-1.0-0.50.00.51.0 6 nN 9 nN 12 nN 15 nNI (nA)V (V)
(a)
(b)
Figure 5
23
2.83.03.23.43.63.84.0110100100010000 6 nN 9 nN 12 nNI (pA)1000/T2.83.03.23.43.63.84.0100100010000 6 nN 9 nN 12 nNI (pA)1000/T(a)
(b)
Figure 6
24
681012101001000 288 K 308 K 318 K 338 KI(pA)Tip Force (nN)681012100100010000100000 288 K 338 KI (pA)Force (nN)Supporting Information
Temperature and force dependence of nanoscale
electron transport via the Cu protein Azurin
Wenjie Li,† Lior Sepunaru,†, ‡ Nadav Amdursky,†,‡ Sidney R. Cohen&, Israel Pecht,§ Mordechai
Sheves,*,‡ and David Cahen*,†
†Department of Materials & Interfaces, ‡Department of Organic Chemistry, &Chemical Research
Support and §Department of Immunology, Weizmann Institute of Science, Rehovot 76100, Israel
*Corresponding authors:
David Cahen
e-mail: [email protected]
Mordechai Sheves
e-mail: [email protected]
25
AFM imaging
The topography of the self-assembled monolayer of proteins was characterized by AFM in semi-
contact mode under N2 purge. During a tapping cycle the tip spends only a small fraction of the
cycle in repulsive contact with the surface. Therefore, the averaged force per cycle is quite small, as
is the total energy imparted. An additional advantage to the semi-contact mode for obtaining clear
images is that the tip detaches from the surface many times at each pixel, so that there is no shear
force applied, which could lead to sample damage.
The much smaller force that can be used in the tapping mode than in CP-AFM is due to the different
forces that are required for reproducible, stable mechanical or electronic contacts. The differences
arise, because different physical interactions are involved. Quantitative calculation of the applied
force is a complicated function of surface and probe properties, as well as specific operating
conditions. Notwithstanding this uncertainty, with proper tuning of operating conditions, the forces
applied during topographical imaging are at least an order of magnitude smaller than those used for
CP-AFM scans.1
Contact mode scratching procedure
The scratching procedure was performed in a Solver P47 SPM system (ND-MDT, Zelenograd
Russia). The operation mode of the AFM was temporarily switched to the contact mode with a
programmed script and a 1x1 µm2 square area was scanned with a large tip force (160 nN). The
applied force is sufficiently large to scratch away the monolayer , but not sufficient to scratch the
substrate surface. Following the contact mode scratching procedure we switched back to semi-
contact mode to re-scan over a larger area, centered around the resulting hollow (Figure S1 a).
26
(a)
(b)
1µm
Figure S1. (a) AFM topography and (b) line profile of patterned Az monolayer from which part was
removed by the AFM probe, enabling an estimate of the monolayer thickness (b). The island at the
top center of (a) is the deposit of the proteins which were removed from the hollow in the image.
27
01230246 Height (nm)Line profilem(a)
(b)
Figure S2. (a) Raw data of I-V curves for junctions with holo-Az (black squares) and apo-Az (red
dots), measured at room temperature. (b) The I-V curves for the junctions with apo-Az of (a), but
with expanded current scale. Tip force: 6 nN.
Reference
1. Garcia, R., and Perez, R., Dynamic atomic force microscopy methods, Surf. Sci. Rep., 2002,
Vol, 47, pp 197-301
28
-1.0-0.50.00.51.0-1.0-0.50.00.51.0 I (nA)V (V) -1.0-0.50.00.51.0-0.10-0.050.000.050.10 I (nA)V (V) |
1209.1263 | 1 | 1209 | 2012-09-06T11:56:15 | On the Lubensky-Nelson model of polymer translocation through nanopores | [
"physics.bio-ph"
] | We revisit the one-dimensional stochastic model of Lubensky and Nelson [Biophys. J 77, 1824 (1999)] for the electrically driven translocation of polynucleotides through alpha-hemolysin pores. We show that the model correctly describes two further important properties of the experimentally observed translocation time distributions, namely their spread (width) and their exponential decay. The resulting overall agreement between theoretical and experimental translocation time distributions is thus very good. | physics.bio-ph | physics |
On the Lubensky-Nelson model of polymer
translocation through nanopores
Peter Reimann1
Andreas Meyer
Sebastian Getfert
Universitat Bielefeld,
Fakultat fur Physik, 33615 Bielefeld, Germany
1Correspondence: [email protected]
Abstract
We revisit the one-dimensional stochastic model of Lubensky and Nelson
[Biophys. J 77, 1824 (1999)] for the electrically driven translocation of
polynucleotides through α-hemolysin pores. We show that the model cor-
rectly describes two further important properties of the experimentally ob-
served translocation time distributions, namely their spread (width) and
their exponential decay. The resulting overall agreement between theoreti-
cal and experimental translocation time distributions is thus very good.
Key words: Nanopores; translocation; Stochastic modeling; Brownian
motion; α-Hemolysin
Polymer translocation through nanopores
2
Introduction
The translocation of biopolymers such as DNA, RNA, or polypeptides through
protein pores plays a key role in various cellular processes (1). Apart from
these biological systems, also artificial, so-called solid-state nanopores have
recently attracted a lot of attention due to their promising potential as a
new generation of fast and cheap DNA sequencing devices and other medi-
cal diagnostics applications (2). To achieve such goals, many experimental
problems still have to be solved, and also the theoretical understanding and
control of those fundamental transport processes needs substantial further
development.
Here, we reconsider one of the earliest and best established theoreti-
cal models in this context, originally introduced in 1999 by Lubensky and
Nelson (3), and further studied and developed in numerous subsequent
works, see e.g. (4 -- 11). Motivated by the seminal experiments on polynu-
cleotide translocation through an α-hemolysin pore by Kasianowicz et al.
(12), Lubensky and Nelson proposed a theoretical description in terms of
a one-dimensional stochastic model dynamics in a tilted periodic potential
(3). While many features of the experimentally observed translocation time
statistics could indeed be explained remarkably well by their simple model,
the theoretical spread of the translocation times underestimated the exper-
imental one by about two orders of magnitude (3). This discrepancy was
pointed out once again in the review paper (1), but to the best of our knowl-
edge has remained a tacitly ignored problem of such a model ever since. To
resolve this problem is a first main issue of our present work.
Since the quantitative details and sometimes even the qualitative findings
notably depend on the considered pores and polymers, we follow Lubensky
and Nelson in specifically focusing on the experimentally best studied case of
the α-hemolysin protein pore and polynucleotides of single stranded DNA or
RNA. In particular, for this system the translocation time distributions are
quantitatively quite well documented, not only with respect to their above
mentioned spread but also with respect to their decay for large times (13 --
16). The second main point of our paper is that the model of Lubensky and
Nelson also correctly reproduces the experimentally observed exponential
decay.
The overall result is a very good comparison of the complete theoret-
ically predicted translocation time distributions with experimentally ob-
served data sets.
Polymer translocation through nanopores
3
Experimental System
The basic experimental set up is illustrated in Fig. 1. Charged, single-
stranded polynucleotides (DNA or RNA) in aqueous solution are exposed
via electrodes to an externally applied voltage. Two fluid compartments
are separated by a phospholipid membrane and are connected by a sin-
gle α-hemolysin protein pore. Since the phospholipid membrane is non-
conducting, practically the entire voltage drop occurs within the pore and its
immediate neighborhood. Whenever a polynucleotide diffusively approaches
the pore from the "upper" side in Fig. 1, the electrical forces direct it into
the pore and drive it to the other side of the membrane. Every such translo-
cation process is experimentally observable as a reduction of the electrical
current through the pore. Even though the polynucleotides are (practically)
identical, the durations of the current blockades exhibit quite significant
statistical variations. The main theoretical task is to qualitatively explain
and quantitatively model the experimentally observed translocation time
distributions. For further details, see, e.g., (1 -- 16).
Model
According to Lubensky and Nelson (3), the polymer translocation process
is modeled by means of a single dynamical state variable x(t) (slow/relevant
collective coordinate), defined as the contour length of that part of the
polymer chain which already has passed through the pore until time t. In
particular, hydrodynamic (dissipative) and steric (entropic) effects of the
chain segments outside the pore and its immediate neighborhood are con-
sidered as negligible. The most immediate justification of this approxima-
tion is that otherwise a disagreement with the experimentally observed lin-
ear dependence of the mean translocation time upon the polymer length
(1, 11, 12, 14, 15, 17, 18) seems practically unavoidable (19)1. This general
fact is nicely illustrated e.g. in Ref. (22) by means of a model very similar
in spirit to the one by Lubensky and Nelson, but in addition taking into
account the polymer degrees of freedom far from the pore region within an
approximative, accompanying equilibrium description originally due to (23).
The state variable x(t) is subjected to several kinds of forces, most no-
tably due to the externally applied voltage and the electrostatic, mechanical,
1While this experimental finding is beyond any doubt in the case of polynucleotide
translocation through α-hemolysin pores (1, 11, 12, 14, 15, 17, 18), the corresponding
results in the case of solid-state nanopores (18) are contradictory (20, 21). For this reason
the Lubensky-Nelson model may be inappropriate in such a case.
Polymer translocation through nanopores
4
and chemical interaction of the polymer with the pore walls, but also due to
entropic forces within the pore and its immediate neighborhood, generated
by the numerous microscopic degrees of freedom of the ambient solvent, the
pore, and the polymer itself. All those forces can be considered to arise as
minus the derivative of a free-energy type potential of mean force Φ(x). The
remaining effects of the fast molecular degrees of freedom are approximately
modeled as friction (dissipation) and noise (thermal fluctuations), while in-
ertia effects are usually negligible on those small lengths and velocity scales.
Altogether, we thus arrive at an overdamped Langevin dynamics of the well
established form (24, 25)
η x(t) = −Φ′(x(t)) +p2ηkT ξ(t) ,
(1)
where ξ(t) is a delta-correlated Gaussian white noise, η is the friction coef-
ficient, and kT the thermal energy.
In the simplest case of a homopolymer, the force −Φ′(x) remains invari-
ant when the entire polymer is translocated by the length a of one monomer,
i.e. Φ′(x + a) = Φ′(x) for all x. As a consequence, Φ(x) must be a tilted pe-
riodic potential, consisting of a strictly a-periodic part U (x) and a constant
"tilting" force F ,
Φ(x) = U (x) − F x .
(2)
Advancing the polymer by one monomer length a changes its (free) en-
ergy by Φ(x+a)−Φ(x) = −aF according to Eq. 2. Following Lubensky and
Nelson (3), the same change of state is obtained by moving one monomer
from one to the other end of the polymer chain. The energy required for
such a move is qV , where q is the charge of a monomer and V the externally
applied voltage (with sign convention as indicated in Fig. 1). We thus can
conclude that (3)
We remark that the the nominal charge per nucleotide is equal to
a F = −q V .
qe = −1.602... · 10−19 C (electron charge).
(3)
(4)
However, it is by now well established (26 -- 31) that due to various electroki-
netic effects of the ambient ionic solution and the pore (screening, electroos-
mosis, electrophoresis, polarization and field confinement mechanisms), the
relevant effective charge q in Eq. 3 is reduced by roughly a factor of 10
compared to the nominal (bare) charge from Eq. 4, i.e.
q ≈ 0.1 qe .
(5)
Polymer translocation through nanopores
5
Due to the above mentioned divers effects which contribute to the charge
renormalization, the exact value of q depends, among others, on temperature
and ion concentrations, but also on the specific monomer (nucleotide) of
which the polynucleotide is composed.
Under the assumption that V , a, and q are (approximately) known,
the force F in Eq. 2 is thus fixed through Eq. 3. Much more difficult to
theoretically estimate from first principles are the friction coefficient η in Eq.
1 and the periodic potential U (x) in Eq. 2. They may thus be considered
as a model parameter and a model function, respectively, which remain to
be determined by experimental means.
Velocity and Diffusion
As a first quantity of interest we consider the average translocation velocity
v of the polymer through the pore. Focusing on not too short polymers,
"boundary-effects" while the polymer enters and exits the pore are negligible
and v follows as the time- and ensemble-averaged velocity x(t) from the
model in Eq. 1 with an infinitely extended periodic potential U (x) in Eq. 2.
The analytical solution of this problem goes back to Stratonovich (32) and
has subsequently been rederived many times, see e.g. chapter 11 in (24).
Adopting the notation from (33, 34), this solution takes the form
v =
kT
aη
1 − e−aF/kT
R a
dx
a I(x)
0
,
where we have introduced
I(x) = eΦ(x)/kT Z x
x−a
dy
a
e−Φ(y)/kT .
(6)
(7)
A further quantity of interest is the random spread of the transloca-
tion velocity (and thus of the translocation time) about its mean value v,
quantified by the diffusion coefficient
D = lim
t→∞
h[x(t) − x(0) − vt]2i
2t
(8)
where h·i indicates an average over the noise ξ(t) in Eq. 1 and over the
initial positions x(0).
Similarly as for the velocity v, the analytical result for the diffusion
coefficient in a tilted periodic potential, Eqs. 1, 2, has been independently
obtained several times. To the best of our knowledge, the first closed, exact
Polymer translocation through nanopores
6
expression for D is buried in the paper (35). For the second time, the same
problem was solved again by Lubensky and Nelson, see Appendix B in (3).
Further rediscoveries are due to (36) and (33, 34). While all those results
are of course equivalent, the actual formulae for D are quite different and,
with the exception of (33, 34), also quite involved. For this reason, the one
from (33, 34) is most common, reading
D =
where we have introduced
kT
0
η R a
dx
a I 2(x) J(x)
a I(x)(cid:3)3
(cid:2)R a
dx
0
,
J(x) = e−Φ(x)/kT Z x+a
x
dy
a
eΦ(y)/kT .
(9)
(10)
The main quantity of interest later on will be the dimensionless ratio
av/D (cf. Section "Spread of Translocation Times" and Ref. (3)), given
according to Eqs. 6 and 9 by
av
D
= (cid:2)1 − e−aF/kT(cid:3) (cid:2)R a
0
dx
a I 2(x) J(x)
dx
a I(x)(cid:3)2
.
(11)
R a
0
A first main result of our paper consists in the observation that the
leading order behavior of Eq. 11 for small values of aF/kT takes the form
av
D ≃
aF
kT
,
(12)
independently of any further details of the periodic potential U (x). In the
simplest case, this result follows by expanding the left bracket in the nu-
merator of Eq. 11 to first (=leading) order in aF/kT and evaluating all the
remaining integrals for aF/kT = 0 (leading=zeroth order in aF/kT ).
In
doing so, the integrals in Eqs. 7 and 10 become x-independent and, as a
consequence, the denominator in Eq. 11 becomes equal to the right bracket
in the numerator. An analogous (but more tedious) expansion resulting in
12 is also contained in (3). Both expansions, however, become question-
able in the weak noise limit (small thermal energy kT ), since the expansion
coefficients in general will inherit exponentially large values from the inte-
grands exp{±[U (x)−U (y)]/kT} contributing via Eqs. 2, 7, 8 to the multiple
integrals in 11.
Our first remark is that Eq. 12 in fact still remains valid for asymptoti-
cally weak noise (kT → 0) and not too large F -values, so that the dynamics
Polymer translocation through nanopores
7
in Eqs. 1, 2 is governed by rare, thermally activated transitions between
metastable states:
In this case, the dynamics can be approximately de-
scribed by a one-dimensional random walk between discrete sites at distance
a with certain forward and backward hopping rates r+ and r−, respectively.
As a consequence (25), one obtains v = a(r+ − r−) and D = a2(r+ + r−)/2.
Furthermore, detailed balance symmetry (25) implies for the forward and
backward rates the relation r+/r− = exp{aF/kT}. We thus obtain the
asymptotically exact result
av
D
= 2
1 − e−aF/kT
1 + e−aF/kT
= 2 tanh(aF/2kT ) ,
(13)
independently of any further details of U (x). In particular, for small aF/kT
one readily recovers Eq. 12.
Our second remark is that both for asymptotically large F and for
asymptotically large kT , the effects of the periodic potential U (x) in Eq.
2 become negligible (33, 34) with the consequence that the exact expression
in Eq. 11 approaches once again the asymptotics of Eq. 12. The same
conclusion is of course recovered if the variations of the periodic potential
U (x) itself become negligibly small.
Our last remark is that the diffusion coefficient D as a function of the
tilt F develops an arbitrarily pronounced maximum for sufficiently small
kT , see (33, 34). Nevertheless, closer inspection along the lines of (33, 34)
shows that the ratio D/av remains a strictly decreasing function of F within
the neighborhood of the maximum of D.
Translocation time distribution and exponential de-
cay
A "successful translocation event" starts when the polymer enters the pore
from one side and ends when it exits at the other side.
In contrast, if
the polymer exits at the same side as it entered, we are dealing with an
"unsuccessful translocation attempt". Following Lubensky and Nelson, we
ignore unsuccessful attempts and henceforth only consider the successful
events. Regarding the experimental identification of such events see e.g.
(16). The statistical distribution of their duration is the quantity of foremost
interest from now on.
A main achievement of Lubensky and Nelson's work (3) is an analytical
approximation for the distribution (probability density) ψ(t) of transloca-
tion times.
In fact, the approximation becomes asymptotically exact for
Polymer translocation through nanopores
sufficiently large numbers N of monomers, i.e.
N = L/a ≫ 1 ,
8
(14)
where a and L denote the lengths of one monomer and of the entire polymer,
respectively.
Remarkably enough, but also quite plausible at second glance, the only
parameters entering the translocation time distribution ψ(t) are the polymer
length L and the velocity v and diffusion coefficient D of the corresponding,
infinitely extended dynamics. Furthermore, it is convenient to employ the
rescaled, dimensionless time
τ =
v t
L
and the dimensionless auxiliary parameter
κ =
4 D
v L
.
(15)
(16)
(17)
(18)
Referring to Appendix A of (3) for the detailed calculations, the final ana-
lytical expression provided by eq. (A6) in (3) and can be rewritten in the
form
ψ(t) =
c
τ 3/2 Xn=1,3,5,...
n2
2
κ τ − 1
κτ o
κ + (τ −n)2
expn 2(n−1)
where the normalization constan c is given by2
c =
v
L r κ
π
[1 − e−4/κ] .
In other words, the translocation time distribution ψ(t) actually does not
depend separately on all three parameters L, v, and D, but only on the two
specific combinations v/L and κ = 4D/vL.
The behavior of Eq. 17 for large t is not obvious at all. In particular, for
large τ the summands on the right hand side are negative up to quite large
n-values, while those for even larger n are positive. In view of the property
ψ(t) ≥ 0 and the expected asymptotics ψ(t) → 0 for t → ∞, there must
be a very fragile cancellation of positive and negative summands. On the
other hand, we observe that, according to eq. (A2) in (3), an asymptotically
exponential decay of the right hand side in Eq. 17 may be conjectured with
a decay rate
λ = (π/2)2κ + 1/κ .
(19)
2There is a typo on the right hand side of eq. (A6) in (3): the first factor 2 should be
replaced by 1/2
Polymer translocation through nanopores
9
In view of this surmise, we thus may rewrite Eq. 17 in the following equiv-
alent form
ψ(t) = c′ g(κ τ ) e−λτ ,
(20)
where c′ is another normalization constant, and where the auxiliary function
g(x) is defined as follows
g(x) = 0.45731
exπ2/4
x3/2 Xn=1,3,5,...
n2
x − 1
2
en2/x
.
(21)
As can be seen from Fig. 2, the function g(x) converges towards a finite
limit for x → ∞ and the specific numerical factor 0.45731 in Eq. 21 has
been chosen so that this limit is (practically) unity.
This brings us to the next main result of our paper: According to Eq.
20 and Fig. 2, the distribution of translocation times ψ(t) predicted by the
model of Lubensky and Nelson exhibits an exponential decay for large times,
in agreement with the experimental findings from (13 -- 16).
We finally remark that equation 20 together with Fig. 2 and Eqs. 15,
16, 19 provide a quite detailed qualitative picture of the translocation time
distribution ψ(t), and how it depends on the parameters v/L and κ: Initially,
ψ(t) remains close to zero, then increases quite steeply up to a maximum
at tmax = τmaxL/v, where τmax solves κ g′(κτmax) = λ g(κτmax), and finally
approaches an exponential decay.
Numerically, for any given set of parameters v/L and κ the quantitative
evaluation of ψ(t) according to Eqs. 15-18 is straightforward. In particu-
lar, for κ → 0 (vanishing thermal fluctuations) one recovers the expected
deterministic limit ψ(t) → δ(t − L/v). Typical examples will be presented
in Section "Comparison with Experiments" below.
Spread of translocation times
Concerning a quantitative comparison between the theoretical prediction
from Eq. 17 and experimental data, we first observe that there are two
fit parameters: One (namely L/v in Eq. 15) amounts to a quite trivial
rescaling of time, which can be readily fixed e.g. by fitting the theoretical
peak of ψ(t) to the experimentally observed one. The remaining, second
fit parameter is κ from Eq. 16, which can be determined by the following
very convenient procedure originally due to Lubensky and Nelson (3): In
a first step, the peak position tmax of the experimentally observed ψ(t) is
Polymer translocation through nanopores
10
determined, formally defined via ψ′(tmax) = 0. Then the two times tL and
tR to the left (L) and to the right (R) of tmax are determined according to
ψ(tL,R) = e−1/2 ψ(tmax) ≃ 0.606 · ψ(tmax) .
(22)
In other words, tR − tL quantifies the width of the experimentally observed
peak of ψ(t). Observing that the ratio (tR − tL)/tmax is independent of the
chosen units of time, we can readily recast the approximative relation from
Fig. 3 of Lubensky and Nelson's work (3) into the form (tR − tL)/tmax ≃
√2κ. In fact, this approximation becomes exact for asymptotically small
(tR − tL)/tmax and remains quite accurate as long as (tR − tL)/tmax ≤ 1. In
other words, we obtain the quite accurate estimate
1
tmax (cid:19)2
2 (cid:18) tR − tL
κ ≃
provided
tR − tL
tmax ≤ 1 .
(23)
On the other hand, combining Eqs. 3, 12, 14, and 16 yields the relation
kT
(−q) V ≃
κ N
4
provided
κ N
4 ≥ 1 .
(24)
From the experimental data in Fig. 2 of Kasianowicz et al. (12) one
readily reads off (tR − tL)/tmax ≃ 1 (3), implying with Eq. 23 that κ ≃ 0.5.
Taking into account that N = 210 in the experiments from (12), the right
hand side of Eq. 24 amounts to κN/4 ≃ 26. Hence the condition κN/4 ≥ 1 is
fulfilled and we can apply Eq. 24 to conclude that kT /qV ≃ 26. On the other
hand, using T ≃ 293 K (room temperature), V = 120 mV (experimental
voltage from (12)), and the nominal charge q = qe ≃ −1.6 · 10−19 C per
nucleotide from (3), we obtain the result kT /(−q)V ≃ 0.21.
In view of
the discprepancy with the relation kT /(−q)V ≃ 26 following from Eq. 24,
Lubensky and Nelson concluded that their model, Eq. 1, which implied Eq.
24, was inconsistent with the experimental facts.
In the following, we argue that this conclusion is not tenable. Rather,
the prediction from Eq. 24 of the model, Eq. 1, agrees quite well with the
experimental findings. To this end, we first evaluate the right hand side
of Eq. 24 by means of the results from several more recent, and therefore
possibly more accurate experiments than in the original work of Kasianowicz
et al. (12): From the two data sets of Meller at al. displayed in Fig. 2 of
their work (13) (see also Fig. 3 in (14) and Fig. 6 in (15)) one can infer
with Eq. 23 that κ ≃ 0.09 and κ ≃ 0.11, respectively (see also next Section).
With N = 100 (13) it follows that κN/4 ≃ 2.3 and κN/4 ≃ 2.8, respectively.
Likewise, one can infer from the data in Fig. 2 of Bates et al. (37) that
Polymer translocation through nanopores
11
κ ≃ 0.17 and hence with N = 60 that κN/4 ≃ 2.6. Finally the data from
Fig. 5 of Butler et al. (16) imply that κ ≃ 0.19 and with N = 50 that
κN/4 ≃ 2.4.
We remark that in all those experiments homopolymers have been used
and that we do not know of any further data of this kind (histograms of
translocation durations for homopolymers) in the literature. For the sake
of completeness, we may also include here the data for heteropolymers with
N = 92 from Fig. 2c in Maglia et al. (38), yielding κ = 0.13 and hence
κN/4 ≃ 3.0.
Turning to the quantity kT /(−q)V appearing on the left hand side of
Eq. 24, we observe that the voltage V = 120 mV, and the temperature
T ≃ 293 K was (approximately) the same in all the experiments from (12 --
16, 37, 38). Using the nominal charge q = qe per nucleotide from (3),
one thus recovers the same result kT /(−q)V ≃ 0.21 as before, and hence
Eq. 24 still seems to be violated. However, according to (26 -- 31) a much
more realistic estimate follows by employing the appropriately renormalized
effective charge from Eq. 5, namely kT /(−q)V ≃ 2.1. As a consequence,
the theoretically predicted relation in Eq. 24 is satisfied quite well by all
the more recent pertinent experiments (13 -- 16, 37, 38).
Taking the model of Lubensky and Nelson and thus Eq. 24 for granted,
the above findings for κN/4 may now in turn be used to estimate the renor-
malized charge q more accurately than in Eq. 5. Since temperature, voltage,
buffer etc. were almost the same in all cases, the resulting differences in q
must be mainly due to the different nucleotides (see also below Eq. 5).
With respect to the earlier experiment by Kasianowicz et al.
(12), a
further reduction of the effective charge by another factor of 10 would be
a possible, though not very satisfying, explanation (see also next Section).
A more likely reason seems to be connected with the considerably larger
κN/4-value compared to the more recent experiments. In other words, the
experimentally observed spread of the translocation times is unusually large.
Indeed, one generally expects that the experimentally observed spreads of
the translocation times still somewhat overestimate the purely diffusive ef-
fects accounted for in the theory. One possible reason of why the observed
spread was particularly large in (12) may be that the data analysis and
pre-processing according to their detailed current blockade signatures was
not yet as sophisticated as e.g.
in (13 -- 16). Another possible factor is the
smaller measurement bandwidth of about 24 KHz for the experiments (12),
compared to 100 KHz for (13 -- 15, 37) and 50 KHz for (16). As will be ar-
gued in the next section, the most plausible explanation is a relatively large
spread of the polynucleotide lengths of the samples used by Kasianowicz et
Polymer translocation through nanopores
12
al. (12).
Comparison with experiments
The purpose of this Section is a comparison between the complete transloca-
tion time distributions for several of the experiments already considered in
the previous Section and the corresponding theoretical distributions. Quite
surprisingly, such a detailed quantitative comparison does not seem to exist
in the previous literature known to us.
As a first example, Fig. 3 presents the experimental data reported by
Butler et al. (16) for single-stranded RNA rC50 polynucleotides (i.e. N = 50
in Eq. 14). The theoretical translocation time distribution has been ob-
tained as described in the previous Section: First, one readily reads off from
the experimental data points that (tR − tL)/tmax ≃ 0.61, yielding with Eq.
23 the estimate κ ≃ 0.19. Then, the time-scale parameter L/v in Eq. 15 is
adapted so as to optimally fit the peak position of the experimental data,
resulting in the estimate L/v ≃ 0.2 ms. According to Fig. 3, the theoretical
ψ(t) obtained in this way agrees very well with the experimental findings
with the exception of the times t smaller than about 0.1 ms. In fact, the
corresponding experimental data were denoted in Ref. (16) as "ambiguous
signals", possibly caused by "retraction of the threaded configuration back
into the vestibule configuration, very rapid translocation, or translocation of
short polynucleotide fragments. Due to the ambiguity in the interpretation
of these short Deep states, we only designated Deep states with durations
longer than the minimum as translocations." This ambiguity in the inter-
pretation of the experimental current blockades ("Deep states") seems to us
a sufficiently convincing explanation of the deviations from the theoretical
curve at small times t. An apparently rather similar situation for small t
has in fact been discussed already in (3, 12, 42). In view of this ambiguity
for t < 0.1 ms we fitted in Fig. 3 the scaling factor c of the theoretical curve
17 as well as possible to the experimental data for t > 0.1 ms rather than
normalizing it according to 18.
As a second example, Fig. 4 shows the data of Bates et al. (37) for single-
stranded DNA dA60 polynucleotides (i.e. N = 60 in Eq. 14). Theoretically,
we proceeded as before with (tR − tL)/tmax ≃ 0.58, κ ≃ 0.17, and L/v ≃
0.42 ms. Again, the overall agreement between theory and experiment is
very nice with the exception of large times t3.
3In fact, the authors of Ref. (37) mention that the tail of the experimentally observed
distribution extends to even much longer times without providing the actual data.
Polymer translocation through nanopores
13
As pointed out by an anonymous referee, the latter disagreement can be
naturally explained by the well-known directionality of the polynucleotide's
sugar-phosphate backbone, resulting in two different translocation time dis-
tributions depending on whether the DNA enters the pore with its 3' or 5'
end first (3, 12, 16, 39 -- 45). Denoting the probabilities of 3' and 5' entries
by p and 1 − p and the concomitant two distributions by ψ1(t) and ψ2(t),
the total (experimentally observed) distributions is given by
ψ(t) = pψ1(t) + (1 − p)ψ2(t) .
(25)
Within the model of Lubensky and Nelson, ψ1(t) and ψ2(t) are both of the
form 17. For both of them, the parameter κ must be the same according
to Eqs. 3, 12, 16 under the plausible assumption that the effective charge
q is (approximately) the same for both DNA orientations (see also below
Eq. 5). On the other hand, there may in general be two different time
scales L/v1 and L/v2, and likewise for τ in Eq. 15. Along these lines,
the best fit to the experimental data of Bates et al. (37) was obtained for
κ ≃ 0.18, L/v1 ≃ 0.42 ms, L/v2 ≃ 0.82 ms, and p ≃ 0.81. According to Fig.
5, the resulting agreement between experiment and theory is indeed very
good. We remark that while quantitative experimental estimates for the
two "extra parameters" p and v1/v2 of the extended theory 25 do not seem
available, the "event diagrams" for dA50 in Fig. 3 of the paper (16) and
for dA100 in Fig. 2 of the paper (40) qualitatively compare very favorably
with our above findings p ≃ 0.81 and v1/v2 ≃ 2 for dA60. We finally remark
that also in the works (13, 40) two "groups" of dA100 translocation events
were identified, "group 1" containing about 80% of the events (at 20◦ C)
(13), and that the two "groups" can be attributed to the two different DNA
orientations (40, 42).
As expected, the already very good agreement in Fig. 3 with the rC50
data by Butler et al. (16) could not be notably improved any more by means
of the extended theory from Eq. 25. This is consistent with the "event
diagram" for rC50 in Fig. 3 of the paper (16), evidencing that while the
RNA may indeed again exhibit two different orientations, the translocation
time distributions happen to be very similar for both of them.
We finally return to the experiment of Kasianowicz et al. (12), using
poly[U] samples with a nominal length of N = 210 nucleotides. In a later
work (39), the same group used a poly[U] sample whose length distribu-
tion was specified as N = 150 ± 50 nucleotides. According to a personal
communication by one of the authors (D. Branton, Harvard University), a
comparable polydispersity of N = 210± 70 may thus be considered as quite
Polymer translocation through nanopores
14
plausible also in the earlier work (12). Theoretically, we took into account
this fact by including on the right hand side of Eq. 25 an integral over a
Gaussian length distribution. Formally, this is achieved by replacing L in
the definitions 15 and 16 by z L, where z > 0 is Gaussian distributed4 with
average 1 and standard deviation 1/3, and likewise for the two time scales
L/v1, L/v2 entering into Eq. 25. Fig. 6 shows our best fit to the experimen-
tal data of Kasianowicz et al. (12), obtained for κ ≃ 0.1, L/v1 ≃ 0.31 ms,
L/v2 ≃ 1.3 ms, and p ≃ 0.58. The agreement is obviously very good, except
for the leftmost data point in Fig. 6. In fact, there are additional experi-
mental data points at even smaller t-values with ψ-values beyond the range
displayed in Fig. 6. Similarly as in Fig. 3, they are commonly considered
to be due to polymers that partially entered the channel but then retracted
rather than actually traversed the channel (3, 12, 16, 42). Such events are
not covered by our present theory, explaining the disagreement in Fig. 6
at small t. Accordingly, the experimental data in Fig. 6 have not been
normalized to unity but rather so that the agreement with the theory was
optimal.
We remark that such "unsuccessful translocation attempts" could be
sorted out in the experimental data in Figs. 4 and 5 thanks to their very
specific current blockade signatures (37). Hence the theory explains the data
even at small t. A similar identification of such events in the experiments
by Kasianowicz et al. (12) and by Butler et al. (16) was apparently not
possible.
Returning to Fig. 6, the obtained fit κ = 0.1 implies that κN/4 =
5.25 and thus with Eqs. 24 and 4 that q ≈ 0.04 qe. Such an effective
charge is smaller than the estimate from equation 5 but still reasonably close
to our previously obtained q-values5. We also remark that by assuming a
larger polydispersity of N = 210 ± 95, the data from Fig. 6 could be fitted
practically equally well but now with κ = 0.05 and hence q ≈ 0.08 qe, and
likewise for N = 210 ± 120 and any q > 0.5 qe. In contrast, without any
polydispersity the best fit becomes somewhat worse than in Fig. 6 and
q ≈ 0.015 qe becomes unrealistically small.
Finally, we also fitted the above extended model to the experimental
data sets from Figs. 3 and 5 but for any non-negligible polydispersity this
always resulted in a less good agreement than without any polydispersity.
4Since z ≤ 0 is ruled out, the Gaussian must be truncated and properly renormalized.
5Note that also the "blocking currents" may vary substantially for different nucleotides
and that this may be closely related to variations of q (27).
Polymer translocation through nanopores
15
Discussion
In their seminal work (3), Lubensky and Nelson showed that their simple
one-dimensional stochastic model, Eqs. 1-3, captures many of the exper-
imental observations on polynucleotide translocation through α-hemolysin
nanopores. However, the quantitative behavior of the experimental translo-
cation time distribution of Kasianowicz et al. (12) could not be satisfactorily
explained. Here, we resolved this long standing problem by taking into ac-
count that due to various electrokinetic effects the relevant effective charge
of the nucleotides is substantially reduced compared to their bare (nominal)
charge.
A second main point of our work was to show that the model of Lubensky
and Nelson implies an asymptotically exponential decay of the translocation
time distribution, in agreement with the experimental results from Refs.
(13 -- 16).
Finally, we compared the complete theoretically predicted translocation
time distributions with experimental distributions from the literature. Tak-
ing into account the directionality of polynucleotides, the non-negligible
polydispersity in the experiment by Kasianowicz et al. (12), and the fact
that "unsuccessful translocation attempts" are not covered by the theory,
the model of Lubensky and Nelson explains the experimental observations
remarkably well.
Regarding alternative, more sophisticated descriptions e.g.
in term of
bead-spring models with many degrees of freedom (19), one naturally ex-
pects that -- at least within certain regimes or limits of the various model
parameters -- the main features of the one-dimensional model of Lubensky
and Nelson should be recovered. Particularly important such features are
the experimentally observed linear dependence of the mean translocation
time upon the polymer length (see also Section "Model") and the main
quantitative characteristics of the observed translocation time distributions,
most notably their spread and exponential decay. The essential open ques-
tion with respect to those more sophisticated models is then, whether the
numerous extra degrees of freedom in order to describe the hydrodynamic
and entropic effects outside the immediate pore region still play a significant
role within the above mentioned, relevant model parameter regimes.
-- -- -- -- -- -- -- -- --
We thank D. Branton for a very helpful e-mail exchange regarding Ref. (12).
This work was supported by Deutsche Forschungsgemeinschaft under SFB
613 and RE1344/8-1.
Polymer translocation through nanopores
16
References
1. Meller, A. 2003. Dynamics of polynucleotide transport through
nanometre-scale pores. J. Phys.: Condens. Matter 15:R581-R607.
2. Venkatesan, B. M., and R. Bashir. 2011. Nanopore Sensors for
Nucleic Acid Analysis. Nature Nanotech. 6:615-624.
3. Lubensky, D. K., and D. R. Nelson. 1999. Driven polymer translo-
cation through a narrow pore. Biophys. J. 77:1824-1838.
4. Kafri, Y, D. K. Lubsenky, and D. R. Nelson. 2004. Dynamics of
molecular motors and polymer translocation with sequence het-
erogeneity. Biophys. J. 86:3373-3391.
5. Lua, R. C., and A. Y. Grosberg. 2005. First passage times and
asymmetry of DNA translocation. Phys. Rev. E 72:061918-1-8.
6. Wanunu, M., B. Chakrabarti, J. Math´e, D. R. Nelson, and A.
Meller. 2008. Orientation-dependent interactions of DNA with an
α-hemolysin channel. Phys. Rev. E 77:031904-1-5.
7. Chen, Z., Y. Jiang, D. R. Dunphy, D. P. Adams, C. Hodges, N.
Liu, N. Zhang, G. Xomeritakis, X. Jin, N. R. Aluru, S. J. Gaik,
H. W. Hillhouse, C. J. Brinker. 2010. DNA translocation through
an array of kinked nanopores. Nat. Mat. 9:667-675.
8. Li, J., and D. S. Talaga. 2010. The distribution of DNA translo-
cation times in solid-state nanopores. J. Phys.: Condens. Matter
22:454129-1-8.
9. Lu, B., F. Albertorio, D. P. Hoogerheide, and J. A. Golovchenko.
2011. Origins and consequences of velocity fluctuations during
DNA passage through a nanopore. Biophys. J. 101:70-79.
10. Muthukumar, M. 2010. Theory of capture rate in polymer translo-
cation. J. Chem. Phys. 132:195101-1-10.
11. Wong, C. T. A., and M. Muthukumar. 2010. Polymer transloca-
tion through α-hemolysin pore with tunable polymer-pore elec-
trostatic interaction. J. Chem. Phys. 133:045101-1-12.
12. Kasianowicz, J. J., E. Brandin, D. Branton, and D. W. Deamer.
1996. Characterization of individual polynucleotide molecules us-
ing a membrane channel. PNAS 93:13770-13773.
Polymer translocation through nanopores
17
13. Meller, A., L. Nivon, E. Brandin, J. Golovchenko, and D. Bran-
ton. 2000. Rapid nanopore discrimination between single polynu-
cleotide molecules. PNAS 97:1079-1084.
14. Meller, A., L. Nivon, and D. Branton. 2001. Voltage-driven DNA
translocation through a Nanopore. Phys. Rev. Lett. 86:3435-3438.
15. Meller, A., and D. Branton. 2002. Single molecule measurements
of DNA transport through a nanopore. Electrophoresis 23:2583-
2591.
16. Butler, T. Z., J. H. Gundlach, and M. A. Troll. 2007. Ionic cur-
rent blockades from DNA and RNA molecules in the α-hemolysin
nanopore. Biophys. J. 93:3229-3240.
17. Deamer, D. W., and D. Branton. 2002. Characterization of nucleic
acids by nanopore analysis. Acc. Chem. Res. 35:817-825.
18. Dekker, C. 2007. Solid-state nanopores. Nat. Nanotechnol. 2:209-
215.
19. Milchev, A. 2011. Single-polymer dynamics under constraints:
scaling theory and computer experiment. J. Phys.: Condens. Mat-
ter 23:103101-1-24.
20. Chen, P., J. Gu, E. Brandin, Y.-R. Kim, Q. Wang, and D. Bran-
ton. 2004. Probing single DNA molecule transport using fabri-
cated nanopores. Nano Lett. 4:2293-2298.
21. Storm, A. J., C. Storm, J. Chen, H. Zandbergen, J.-F. Joanny,
and C. Dekker. 2005. Fast DNA translocation through a solid-
state nanopore. Nano Lett. 5:1193-1197.
22. Muthukumar, M. 1999, Polymer translocation through a hole. J.
Chem. Phys. 111:10371-10374.
23. Sung, W,, and Park, P. J. 1996. Polymer Translocation through
a pore in a Membrane. Phys. Rev. Lett. 77:783-786.
24. H. Risken, The Fokker-Planck Equation, Springer, Berlin, 1984.
25. Reimann, P. 2002. Brownian Motors: Noisy Transport far from
Equilibrium. Phys. Rep. 361:57-265.
Polymer translocation through nanopores
18
26. Rabin, Y., and M. Tanaka. 2005. DNA in nanopores: counterion
condensation and coion depletion. Phys. Rev. Lett. 94:148103-1-4.
27. Zhang, J., and B. I. Shlokovskii. 2007. Effective charge and free
energy of DNA inside an ion channel. Phys. Rev. E 75:021906-1-
10.
28. Ghosal, S. 2007. Effect of salt concentration on the electrophoretic
speed of a polyelectrolyte through a nanopore. Phys. Rev. Lett.
98:238104-1-4.
29. Henrickson, S. E., M. Misakian, B. Robertson, and J. J.
Kasianowicz. 2000. Driven DNA transport into an asymmetric
nanometer-scale pore. Phys. Rev. Lett. 85:3057-3060.
30. Luo, K., T. Ala-Nissila, S.-C. Ying, and A. Bhattacharya. 2008.
Sequence dependence of DNA translocation through a nanopore.
Phys. Rev. Lett. 100:058101-1-4.
31. Brun, L., M. Pastoriza-Gallego, G. Oukhaled, J. Math´e, L. Bacri,
L. Auvray, and J. Pelta. 2008. Dynamics of polyelectrolyte trans-
port through a protein channel as a function of applied voltage.
Phys. Rev. Lett. 100:158302-1-4.
32. Stratonovich, R. L. 1958. Oscillator synchronization in the pres-
ence of noise, Radiotekhnika i elektronika 3:497. English trans-
lation in Non-linear transformations of stochastic processes. P. I.
Kuznetsov, R. L Stratonovich, V. I. Tikhonov, editors. Pergamon,
Oxford, 1965.
33. Reimann, P., C. Van den Broeck, H. Linke, P. Hanggi, J. M. Rubi,
and A. P´erez-Madrid. 2001. Giant Acceleration of Free Diffusion
by Use of Tilted Periodic Potentials. Phys. Rev. Lett. 87:010602:1-
4.
34. Reimann, P., C. Van den Broeck, H. Linke, P. Hanggi, J. M. Rubi,
and A. P´erez-Madrid. 2002. Diffusion in tilted periodic potentials:
Enhancement, universality, and scaling. Phys. Rev. E 65:031104-
1-16.
35. Parris, P. E., M. Kus, D. H. Dunlap, and V. M. Kenkre. 1997.
Nonlinear response theory:Transport coefficients for driving fields
of arbitrary magnitude. Phys. Rev. E 56:5295-5305.
Polymer translocation through nanopores
19
36. Lindner, B., M. Kostur, and L. Schimansky-Geier. 2001. Optimal
diffusive transport in a tilted periodic potential. Fluct. Noise Lett.
1:R25-R39.
37. Bates, M., M. Bruns, and A. Meller. 2003. Dynamics of DNA
molecules in a membrane channel probed by active control. Bio-
phys. J 84:2366-2372.
38. Maglia, G., M. R. Restrepo, E. Mikhailova, and H. Bayley.
2008. Enhanced translocation of single DNA molecules through
α-hemolysin nanopores by manipulation of internal charge. PNAS
105:19720-19725.
39. Akeson, M., D. Branton, J. J. Kasianowicz, E. Brandin, and D.
W. Deamer. 1999. Microsecond time-scale discrimination among
polycytidylic acid, polyadenylic acid, and polyuridylic acid as ho-
mopolymers or as segments within single RNA molecules. Bio-
phys. J. 77:3227-3227.
40. Wang, H., J. E. Dunning, A. P.-H. Huang, J. A. Nyamwanda, and
D. Branton. 2004. DNA heterogeneity and phosphorylation un-
veiled by single-molecule electrophoresis. PNAS 101:13472-13477.
41. Math´e, J., A. Aksimentiev, D. R. Nelson, K. Schulten, and A.
Meller. 2005. Orientation discrimination of single stranded DNA
inside the α-hemolysin membrane channel. PNAS 102:12377-
12382.
Butler,
42. Butler, T. Z. , J. H. Gundlach, and M. A. Troll. 2006. Determina-
tion of RNA orientation during translocation through a biological
nanopore. Biophys. J. 90:190-199.
43. Purnell, R. F., K. K. Mehta, and J. J. Schmidt. 2008. Nucleotide
identification and orientation discrimination of DNA homopoly-
mers immobilized in a protein nanopore. Nano Lett. 8:3029-3034.
44. Muzard, J., M. Martinho, J. Math´e, U. Bockelmann, and V. Vias-
noff. 2010. DNA translocation and unzipping through a Nanopore:
some geometrical effects. Biophys. J. 98:2170-2178.
45. Aksimentiev, A. 2010. Deciphering ionic current signatures of
DNA transport through a nanopore. Nanoscale 2:468-483.
Polymer translocation through nanopores
20
Figure Legends
Figure 1
Schematic illustration of the experimental set up: polynucleotide chains
(single-stranded DNA or RNA) translocate through an α-hemolysin protein
pore due to the externally imposed voltage difference between the electrodes.
Figure 2
The function g(x) obtained by numerically evaluating Eq. 21.
Figure 3
Line: Theoretical translocation time distribution ψ(t) according to Eqs. 15-
18 with κ = 0.19 and L/v = 0.2 ms. Symbols: Experimentally observed
translocation times, adopted from Fig. 5 of Butler et al. (16).
Figure 4
Line: Theoretical translocation time distribution ψ(t) according to Eqs. 15-
18 with κ = 0.17 and L/v = 0.42 ms. Symbols: Experimentally observed
translocation times, adopted from Fig. 2 of Bates et al. (37).
Figure 5
Line: Theoretical translocation time distribution ψ(t) according to Eqs. 25
(see also main text for details) with κ = 0.18, L/v1 = 0.42 ms, L/v2 =
0.82 ms, and p = 0.81. Symbols: Experimentally observed translocation
times, adopted from Fig. 2 of Bates et al. (37).
Figure 6
Line: Theoretical translocation time distribution ψ(t), accounting for poly-
dispersity by randomizing Eqs. 25 (see main text) with κ = 0.1, L/v1 =
0.31 ms, L/v2 = 1.3 ms, p = 0.58. Symbols: Experimentally observed
translocation times, adopted from Fig. 2 of Kasianowicz et al. (12).
Polymer translocation through nanopores
21
Figure 1:
Polymer translocation through nanopores
22
1
0.75
)
x
(
g
0.5
0.25
0
0
0.25
0.5
0.75
1
x
Figure 2:
Polymer translocation through nanopores
23
]
s
m
/
1
[
)
t
(
ψ
8
6
4
2
0
0
0.1
0.2
t [ms]
0.3
0.4
Figure 3:
Polymer translocation through nanopores
24
]
s
m
/
1
[
)
t
(
ψ
3
2
1
0
0
0.5
1
t [ms]
Figure 4:
1.5
Polymer translocation through nanopores
25
]
s
m
/
1
[
)
t
(
ψ
3
2
1
0
0
0.5
1
t [ms]
Figure 5:
1.5
Polymer translocation through nanopores
26
]
s
m
/
1
[
)
t
(
ψ
2
1.5
1
0.5
0
0
0.5
1
1.5
t [ms]
2
2.5
Figure 6:
|
1110.6467 | 2 | 1110 | 2011-11-16T04:25:53 | Gene Expression Noise Facilitates Adaptation and Drug Resistance Independently of Mutation | [
"physics.bio-ph",
"q-bio.PE"
] | We show that the effect of stress on the reproductive fitness of noisy cell populations can be modelled as first-passage time problem, and demonstrate that even relatively short-lived fluctuations in gene expression can ensure long-term survival of a drug-resistant population. We examine how this effect contributes to the development of drug-resistant cancer cells, and demonstrate that permanent immunity can arise independently of mutations. | physics.bio-ph | physics |
Gene Expression Noise Facilitates Adaptation and Drug Resistance Independently of
Mutation
Daniel A. Charlebois1,2,∗ Nezar Abdennur2,3, and Mads Kaern1,2,3†
1Department of Physics,
University of Ottawa, 150 Louis Pasteur,
Ottawa, Ontario, K1N 6N5, Canada.
2Ottawa Institute of Systems Biology,
University of Ottawa, 451 Smyth Road,
Ottawa, Ontario, K1H 8M5, Canada.
3Department of Cellular and Molecular Medicine,
University of Ottawa, 451 Smyth Road,
Ottawa, Ontario, K1H 8M5, Canada.
(Dated: October 2, 2018)
We show that the effect of stress on the reproductive fitness of noisy cell populations can be
modelled as first-passage time problem, and demonstrate that even relatively short-lived fluctuations
in gene expression can ensure long-term survival of a drug-resistant population. We examine how this
effect contributes to the development of drug-resistant cancer cells, and demonstrate that permanent
immunity can arise independently of mutations.
Gene expression is a stochastic process that enables
genetically identical cells in the same environment to
exhibit phenotypic variation [1 -- 3]. This noise-induced
non-genetic (epigenetic) variability can be beneficial to
cell populations experiencing acute stress by providing a
temporary basis for natural selection [4 -- 7].
Experimental observations suggest that gene expres-
sion is inherently associated with 'epigenetic memory',
defined by the fluctuation relaxation time of a gene prod-
uct within a cell lineage. In human lung cancer cells, this
relaxation time can be as long as four generations [8].
Brock et al. [9] recently argued that epigenetic mem-
ory might accelerate tumour progression by contribut-
ing to the development of drug-resistant cancer cells. In
this hypothesis, phenotypic variability from the noisy ex-
pression of gene X that confers resistance renders some
cells (and their offspring) temporarily insensitive to the
drug, thereby increasing the probability of acquiring a
mutation conferring permanent immunity. In the present
work, we develop a minimal model to study this phe-
nomenon quantitatively.
To study how gene expression noise impacts the dy-
namics of isogenic cell populations under stress, we define
the reproductive fitness (W ) as the number of offspring
produced in the presence of the stressor (i.e. a drug) rel-
ative to that produced in its absence. For simplicity, we
assume that all cells produce offspring at the same rate
in the absence of the drug, and define the generation
time (tD) as the time it takes for each cell to reproduce
once. We set the generation time as unit time and report
all time-scales relative to tD. We also assume that cells
carry the gene X conferring drug resistance when its ex-
pression level x is sufficiently high, and that this gene is
expressed stochastically in individual cells.
The effects of gene expression noise on populations un-
der stress have previously been analyzed to explain why
certain genes have high expression noise [5 -- 7]. In these
analyses, the dependency between gene expression and
reproductive fitness was defined by the integral
(cid:90)
W (t) =
w(x)px(x, t)dx,
(1)
where px(x, t) is the probability distribution function
(PDF) describing the concentration (x) of the gene prod-
uct across the population, and w(x) is the microscopic
fitness function describing the effect of the drug on the
fitness of cells with a given expression level. The basic
concept is illustrated in Fig. 1(a) using a model where
w(x) is described by the Heaviside step function, such
that cells are unable to reproduce if their expression level
is below a critical value, w(x < xc) = 0, and unaffected
by the drug otherwise, w(x ≥ xc) = 1.
In this case,
previous theoretical work [5 -- 7] concluded that high gene
expression noise is beneficial at high drug-doses, since the
fraction of cells expressing above a reproductive thresh-
old xc increases with the width of the initial expression
distribution [Fig. 1(a)]. However, because px(x, t) is as-
sumed fixed at the time of drug treatment, this conclu-
sion is valid only for instantaneous selection effects. The
analysis of prolonged stress exposure necessitates an ap-
proach where selection, inheritance and gene expression
dynamics all contribute to the evolution of the popula-
tion.
Population survival during prolonged drug exposure is
a first-passage time problem.
In the absence of muta-
tions conferring permanent immunity, cells that survive
the initial selection will eventually succumb to the drug
since they cannot maintain high expression indefinitely.
Consider a sub-population of cells with the same level
of x above xc [Fig. 1(b)]. The time interval in which
a given cell can reproduce is the first-passage (or so-
journ) time tS(x), where the threshold xc represents an
absorbing barrier. Although cells are initially identical,
the expression of the drug-resistance gene evolves differ-
ently in different cells, and the time to reach the repro-
ductive threshold is a random variable described by the
first-passage time distribution pS(x, tS) (Fig. 1(b), In-
set). Since only cells with tS(x) > tD reproduce, w(x) in
Eq. 1 is given by
w(x) =
pS(x, tS)dt(cid:48)
S,
(2)
(cid:90) ∞
tD
and the overall fitness of the population at time t can be
written as
(cid:90) ∞
(cid:18)(cid:90) ∞
xc
tD
W (t) =
(cid:19)
pS(x, tS)dt(cid:48)
S
px(x, t)dx.
(3)
The population fitness in Equation (3) has an explicit
solution only in special cases. Previous analyses [5 -- 7]
circumvented this problem, in part, by focusing on initial
selection effects (t → 0). However, even in this limit, it is
also necessary to assume that all cells above the threshold
contribute to fitness (i.e., w(x) = 1 for x > xc).
To investigate more general cases, we used the
Ornstein-Uhlenbeck (OU) process to model the level of
2
gene expression in individual cells [10]. This process can
be described by the Langevin equation
dx(t)
dt
=
1
τ
(µ − x(t)) + c1/2ξt,
(4)
where c and τ are the diffusion constant and the relax-
ation time, respectively, and ξt is Gaussian white noise
((cid:104)ξt(cid:105) = 0, (cid:104)ξtξt(cid:48)(cid:105) = δ(t − t(cid:48))) [11]. The steady-state PDF
of the OU process is a Gaussian distribution with mean
µ and variance σ2 = cτ /2. Without loss of generality, we
set µ = 0 and use the fluctuation time-scale τ to model
the time-scale of epigenetic memory.
The fluctuation time-scale of gene expression has been
determined experimentally in human lung cancer cells
in terms of the 'mixing time' τm, defined as the lag
where the autocorrelation function has decreased by
50% [8]. The mixing time for the stationary OU pro-
cess is τm = τ ln(2). The measured values of τm varied
between 0.5 to 3.0 generations for different genes, corre-
sponding to values of τ between 0.7 to 4.0 generations
for the OU process.
First, we examined the effect of drug treatment on re-
productive fitness after one generation time when the ab-
sorbing barrier is located at xc = 0.
In this case, the
FIG. 1. Epigenetic effects on a cell population exposed to
stress. (a) Schematic of instantaneous selection effects. (b)
Schematic of generalized model. (c) Reproductive fitness at
the time of first division W (t = tD) after the application of a
stress (at xc = 0) as a function of τ or σ2 for fixed σ2 or τ ,
respectively. Analytical curves (solid lines) were obtained via
numerical solution of Eq. (3). (d) W (t = tD) as a function of
xc for high and low σ2. τ and σ2 are scaled by tD. Dashed
lines represent results obtained from Eq. (1), or equivalently
Eq. (3) in the limit τ → ∞.
FIG. 2. Effect of epigenetic memory τ on drug resistance at
various timescales. (a) Top and bottom plots show population
distributions corresponding respectively to short (τ = 0.5)
and long (τ = 10) epigenetic memory and show the fraction of
drug resistant cells (i.e., cells with x > xc) after acute (t = 0)
and a prolonged (t = 10) drug exposures (single realization of
105 cells). (b) W as a function of t for various values of τ . t,
τ and σ2 are scaled by tD.
first-passage time PDF for x > xc is given by [12]
pS(x, tS) =
exp
x√
2πc
(cid:18)
×
1
τ sinh(tS/τ )
(cid:18)−x2 exp(−tS/τ )
2cτ sinh(tS/τ )
(cid:19)3/2
(cid:19)
+
tS
2τ
.
(5)
We evaluated the effects of varying the time-scale of epi-
genetic memory and the noise amplitude by numerical
integration of Eq. (3), using the steady-state OU distri-
bution to describe the initial gene expression distribution.
Figure 1(c) shows the results for fixed noise (σ2 = 1) and
variable τ , and fixed time-scale (τ = 2) and variable σ2.
The time-scale of epigenetic memory significantly af-
fects 'acute' reproductive fitness, even for very long fluc-
tuation relaxation times. For example, when τ = 20,
W is reduced to 0.4, compared with the value of 0.5 ob-
tained (irrespectively of the noise amplitude) in the per-
manent epigenetic memory limit τ → ∞ [Fig. 1(c)]. For
τ = 2, the reproductive fitness is approximately 0.2, and
the majority of cells starting with x > xc are unable to
maintain above-threshold gene expression long enough to
reproduce. In this case, the acute reproductive fitness re-
mains constant, presumably because changing the noise
amplitude for xc = 0 does not change the fraction of cells
with x > xc.
To examine cases where xc > 0, it is necessary to use
numerical simulations since a general closed-form solu-
tion of the first-passage time PDF is not available. For
this purpose, we employed a population simulation algo-
rithm [13] in which gene expression in each of N individ-
ual cells, xi(t) for i = 1, . . . , N , is obtained by solving
Eq. 4 numerically [14]. In these simulations (20 realiza-
tions of 104 cells unless indicated otherwise), cell divi-
sion occurs when a deterministic cell cycle 'clock', which
is reset at each division, reaches tD. Each cell keeps
track of the time since its birth and can only advance its
clock if they maintain gene expression above the thresh-
old. Moreover, cells where xi(t) ≤ xc are assumed to
be fixed and unable to change their expression level (i.e.,
τ = ∞). Simulations were initiated by assigning, to each
cell, random initial values of gene expression and the cell
cycle clock from the steady-state distribution of the OU
process and a uniform distribution [0 : tD], respectively.
Numerical calculations of fitness for xc > 0 identified τ
as a critical determinant of population survival. Specifi-
cally, the fitness of a population with low gene expression
noise can be greater than that of a population with high
noise if the fluctuation relaxation time is sufficiently long.
We observed this in simulations, shown in Fig. 1(d), with
an increased threshold xc for fixed time-scales (τ = 2 or
τ = 5) and two different fluctuation amplitudes (σ2 = 1
or σ2 = 10). When the two populations had the same
finite value of τ , we observed that increased gene ex-
pression noise always provides a fitness benefit (data not
shown). However, as expected from Fig. 1(c), incorporat-
3
ing stochastic gene expression dynamics (i.e., finite val-
ues of τ ) generally yields a significant reduction in fitness
compared to the asymptotic permanent memory limit.
The magnitude of this reduction is sensitive to both the
value of the threshold xc and the value of τ . This is il-
lustrated in Fig. 1(d) where the fitness of the high noise
population is greater than the low noise population only
when the value of xc is sufficiently high.
In our second case, we analyzed the long-term effects
of varying the time-scale of epigenetic memory on popu-
lation dynamics and reproductive fitness. For simplicity,
we focus on the case where xc = 0 and noise is fixed
(σ2 = 1). Figure 2(a) shows representative gene expres-
sion distributions obtained after 10 generation times for
short- and long-term epigenetic memory. When the fluc-
tuation time-scale is short (τ = 0.5, top panel), the num-
ber of cells that may reproduce (i.e., cells with xi(t) > xc)
is reduced over time since, on average, cells reach the ab-
sorbing barrier faster than they reproduce. Correspond-
ingly, given enough time, the population will go extinct.
This is not the case when memory is long (τ = 10, bottom
panel) and the birth rate exceeds the rate of loss at the
absorbing barrier. In addition, the mode of gene expres-
sion distribution shifts to higher values, in resemblance
of experimental observations [7].
Relatively short-term epigenetic memory can result in
permanent drug resistance even in the absence of mu-
tations. This is illustrated in Fig. 2(b), which shows
how the reproductive fitness of populations with differ-
ent memory time-scales evolve over time. In populations
with long-term memory (e.g., τ = 5, 10, or ∞), the num-
ber of cells that may reproduce increases steady over time
and settles in a steady-state where more than half of
them reproduce every generation time (i.e., W (t) > 0.5).
Importantly, populations with memory at intermediate
time-scales (e.g., τ = 1.5, 2, or 3) may retain long-term
viability and finite rates of reproductive fitness. Because
the simulations involve finite populations, the outcome of
a given realization cannot always be predicted. For ex-
ample, when τ = 1.5, a viable population was observed
to develop in 29% of the simulations while the population
went extinct in the remaining 71% of simulations. While
populations with short memory (e.g., τ = 0.5 or 1) even-
tually go extinct, several cell cycles were needed for the
drug to fully effect all cells.
In the third and final case, we investigated the added
effect of genetic mutations on the development of drug-
resistance. A central element of the Brock et al. hypoth-
esis is that temporary drug resistance due to slow fluc-
tuations in gene expression may contribute to tumour
development by increasing the overall probability that
some cells acquire a mutation conferring permanent im-
munity. To model this scenario, we allowed each cell with
an expression level above xc the chance to mutate once
per generation time. We denote this probability PM . If
a cell acquired the mutation, it and its offspring were
permanently resistant to the drug, and the survival of a
continuously growing population inevitable.
We first investigated the added effect of mutations on
the re-emergence of a cancerous tumour under constant
drug treatment. In these simulations, we chose xc such
that the drug instantaneously removed 95% of the popu-
lation, and measured the time it took for the remaining
cells to double in number. Figure 3(a) shows the depen-
dency of this doubling time on τ when PM is equal to
0.01 and 0.1. These mutation rates are unrealistically
high biologically and were chosen to illustrate the effect
of epigenetic memory in an extreme limit.
As expected, increasing the mutation probability sig-
nificantly reduces the doubling time when the gene ex-
pression fluctuations are short-lived. Unexpected, how-
ever, the value of τ beyond which mutations do not have
an additional effect is remarkably short despite the unre-
alistically high mutation rates. Specifically, the doubling
time is more or less unaffected by PM when τ is roughly
above 4 generations, corresponding to the upper range of
mixing times observed experimentally [8].
We confirmed our results using a semi-realistic model
of gene expression noise [15] where proteins are synthe-
sized in irregular bursts at irregular intervals (Fig. 3(a),
Inset). We also tested the effect of replacing the fitness
threshold with a more realistic sigmoidal fitness function
and found no qualitative difference (data not shown). In
reality, gene expression dynamics may follow more com-
FIG. 3. Effect of τ and PM on cancer cell populations under-
going prolonged drug treatment. a) Effect of PM on doubling
time as a function of τ for an initial population of 1000 non-
mutated cells with gene expression levels above a 95% drug
threshold. Simulation results using burst model of gene ex-
pression [15] shown in the inset. b) Heat maps corresponding
to (a) show probability of remission after 10 generations (100
realizations). PM , τ and σ2 are scaled by tD.
4
plex kinetics than that of a simple mean-reverting pro-
cess due, for example, to multistability and noise-driven
switching [16, 17]. Our simulation results demonstrate
that such complexity is not required for gene expression
noise to have an significant impact on population dynam-
ics under prolonged stress.
We also determined how the probability of remission
depends on the mutation rate, the initial number of can-
cer cells with above-threshold expression, and the time-
scale of gene expression noise. In these simulations, the
cancer is in remission if no cells have above threshold gene
expression and have not acquired a mutation conferring
permanent immunity within 10 generation times. As ex-
pected [Fig. 3b], the probability of remission is greatly
decreased when the number of initial surviving cancer
cells or the mutation rate is increased. Also, when τ
is very short, remission is virtually guaranteed. How-
ever, the probability that a drug-resistant cell population
will emerge can be quite substantial within the experi-
mentally observed range of τ . Even with a relative low
mutation rate (PM = 0.01) and 10 surviving cells, the
probability of remission is only 42% when τ = 4.0.
In summary, we have analyzed the effect of gene ex-
pression noise on the reproductive fitness of isogenic cell
populations under stress as a first-passage time problem.
By explicitly incorporating the 'epigenetic memory' of
this noise (i.e., the fluctuation relaxation time), we have
generalized previous theoretical work that explained the
acute effects of noise amplitude but did not incorporate
gene expression dynamics [5 -- 7]. This generalization is
important for two reasons. First, it has allowed us to
demonstrate using a minimal model that gene expression
noise with biologically realistic time scales has a signif-
icant effect on reproductive fitness under stress and is
a critical determinant of population survival. Second,
it enables theoretical and computational investigations
of experimentally observed phenomena associated with
prolonged stress exposure, including reversible shifts in
gene expression distributions [7], and drug resistance. In
this context, we have demonstrated that the time-scale of
epigenetic memory required to develop a drug-resistant
cell population independently of mutations is compara-
ble to that measured for certain genes in human cancer
cells [8]. Correspondingly, long-term population survival
may not require specialized memory-conferring mecha-
nisms.
It might, for example, be acheived without a
significant fitness cost through bursty gene expression.
An important next step is to confirm our findings us-
ing more realistic models of gene expression incorporat-
ing additional stochastic effects, such as partitioning er-
rors [18], and correspondingly, to employ various analyt-
ical and numerical methods that may permit solution in
these more complex cases (e.g. [19, 20]). We anticipate
that future analysis of such models will provide a deeper
understanding of epigenetic interactions between genes,
drugs and population dynamics.
∗ [email protected]
† [email protected]
[1] M. Kaern, T. Elston, W. Blake,
Nat. Rev. Genet. 6, 451 (2005).
and J. Collins,
[2] A. Eldar and M. Elowitz, Nature 467, 167 (2010).
[3] T. Jia and R. Kulkarni, Phys. Rev. Lett. 105, 018101
(2010).
[4] W. Blake, G. Balazsi, M. Kohanski, F. Isaacs, K. Mur-
and J. Collins,
phy, Y. Kuang, C. Cantor, D. Walt,
Molec. Cell 24, 853 (2006).
[5] D. Fraser and M. Kaern, Molec. Microbiol. 71, 1333
(2009).
[6] Z. Zhang, W. Qian, and J. Zhang, Mol. Syst. Biol. 5,
299 (2009).
[7] D. Zhuravel, D. Fraser, S. St-Pierre, L. Tepliakova,
W. Pang, J. Hasty, and M. Kaern, Syst. Synth. Biol.
(2010), 10.1007/s11693-010-9055-2.
[8] A. Sigal, R. Milo, A. Cohen, N. Geva-Zatorsky, Y. Klein,
Y. Liron, N. Rosenfeld, T. Danon, N. Perzov,
and
U. Alon, Molec. Syst. Biol. (2006), 10.1038/msb4100068.
[9] A. Brock, H. Chang, and S. Huang, Nat. Rev. Genet.
5
10, 336 (2009).
[10] V. Shahrezaei, J. Ollivier, and P. Swain, Mol. Syst. Biol.
4, 196 (2008).
[11] G. Uhlenbeck and L. Ornstein, Phys. Rev. 36, 823 (1930).
[12] M. Wang and G. Uhlenbeck, Rev. Mod. Phys. 17, 323
(1945).
[13] D. Charlebois, J. Intosalmi, D. Fraser, and M. Kaern,
Commun. Comput. Phys. 9, 89 (2011).
[14] D. Gillespie, Phys. Rev. E 54, 2084 (1996).
[15] T. Jia and R. Kulkarni, Phys. Rev. Lett. 106, 058102
(2011).
[16] H. Chang, M. Hemberg, M. Barahona, D. Ingber, and
S. Huang, Nature 453, 544 (2008).
[17] T. Kalmar, C. Lim, P. Hayward, S. Munoz-Descalzo,
J. Nichols, J. Garcia-Ojalvo, and A. Arias, PLoS Biol.
7, e1000149 (2009).
[18] N. Brenner and Y. Shokef, Phys. Rev. Lett. 99, 138102
(2007).
[19] B. Munsky and M. Khammash, IET Syst. Biol. 2, 323
(2008).
[20] M. Stamatakis and K. Zygourakis, J. Theor. Biol. 266,
41 (2010).
|
1005.4919 | 2 | 1005 | 2011-11-08T19:41:51 | Exactly Solvable Model for Helix-Coil-Sheet Transitions in Protein Systems | [
"physics.bio-ph"
] | In view of the important role helix-sheet transitions play in protein aggregation, we introduce a simple model to study secondary structural transitions of helix-coil-sheet systems using a Potts model starting with an effective Hamiltonian. This energy function depends on four parameters that approximately describe entropic and enthalpic contributions to the stability of a polypeptide in helical and sheet conformations. The sheet structures involve long-range interactions between residues which are far in sequence, but are in contact in real space. Such contacts are included in the Hamiltonian. Using standard statistical mechanical techniques, the partition function is solved exactly using transfer matrices. Based on this model, we study thermodynamic properties of polypeptides, including phase transitions between helix, sheet, and coil structures. | physics.bio-ph | physics |
Exactly Solvable Model for Helix-Coil-Sheet Transitions in Protein Systems
John S. Schreck∗ and Jian-Min Yuan†
Department of Physics, Drexel University, Philadelphia, PA 19104
(Dated: February 16, 2018)
In view of the important role helix-sheet transitions play in protein aggregation, we introduce
a simple model to study secondary structural transitions of helix-coil-sheet systems using a Potts
model starting with an effective Hamiltonian. This energy function depends on four parameters
that approximately describe entropic and enthalpic contributions to the stability of a polypeptide
in helical and sheet conformations. The sheet structures involve long-range interactions between
residues which are far in sequence, but are in contact in real space. Such contacts are included
in the Hamiltonian. Using standard statistical mechanical techniques, the partition function is
solved exactly using transfer matrices. Based on this model, we study thermodynamic properties of
polypeptides, including phase transitions between helix, sheet, and coil structures.
PACS numbers: 87.15.Cc, 87.15.A-, 64.60.De
In late the 1950s and early 1960s, Zimm and Bragg
(ZB) and Lifson-Roig (LR) studied helix-coil transitions
of simple models of homopolypeptides by employing rig-
orous statistical methods based on partition functions
and transfer matrices [1].
In the 1970s and 1980s,
these models were extended to include copolymers and
medium-ranged interactions, and were used to character-
ize the experimental results of all amino acids and many
proteins [2]. Because of the close coupling between the
theoretical and experimental studies, ZB, LR, and re-
lated models have stimulated much interest in helix-coil
transitions [3], which is still an active field of research up
to the present time [4, 5]. For reviews, see Ref. [2].
However, conformation changes of polypeptides involv-
ing sheet structures, such as helix-sheet transitions, are
not as well characterized as for helix-coil transitions.
In the late 1970s, using a multi-state model, Tanaka
and Scheraga [6] considered extended and chain-reversal
states in addition to helix-coil transitions.
In Ref. [7],
medium-range interactions were taken into account to
study helices, extended structures, and coils. More re-
cently, Mattice and Scheraga [8], Sun and Doig [9], Hong
and Lei [10], and others have included sheet structures
in statistical models for homo-polypeptides. The diffi-
culties in constructing models for sheets lie primarily in
the interactions between residues that are long-range in
sequence but are close in physical space, and in the rich
variety of structures associated with sheets, turns, and
loops, thus a large number of parameters required for
their description.
In this article, we introduce a sim-
ple statistical mechanical model for helix-coil-sheet tran-
sitions of homo-polypeptides, starting with an effective
Hamiltonian. Instead of an Ising-like model, the treat-
ment is built on a multi-state Potts model, which is ca-
pable of explicitly describing some of the long-range in-
teractions exhibited by sheet structures. The objective is
∗Electronic address: [email protected]
†Electronic address: [email protected]
that this simple model extends the helix-coil treatments
to protein systems with three or more secondary struc-
tures.
An important step in a statistical mechanical approach
like ZB, LR, Ising and Potts models is to construct the
partition function for the system, based on which all ther-
modynamic properties are obtainable. As in ZB and LR
models, partition functions factorize in terms of transfer
matrices. However, ZB or LR theories start with a com-
binatorial partition function without defining an effective
Hamiltonian. More generally, if an energy function H(i)
is defined, where i = (i1, . . . , in) and in is the micro-state
of the nth residue which could occupy one of q possible
states (conformations) labeled as {1, 2, . . . , q}, the par-
tition function for a system of N residues with periodic
boundary conditions reduces to
q
q
q
ZN =
X
X
· · ·
X
e−βH(i) = T r (cid:0)T N(cid:1)
(1)
i1=1
i2=1
iN =1
where β = (kBT )−1, kB is Boltzmann's constant, and
T r is the matrix trace operation. The dimension of a
transfer matrix in a one-dimensional (1D) Ising model is
2 × 2 and for a q-state Potts model, the dimension of a
transfer matrix is q×q. For Potts models with long-range
interactions of range L along a 1D chain, as Glumac and
Uzelac [11] showed in their formulation, the dimension
of a transfer matrix becomes qL × qL. Eq. (1) may be
further simplified by diagonalizing the transfer matrix T .
More recently, Hamiltonians of polypeptide chains
have been described using a variety of Ising-like models
[4, 12, 13] and Potts models [14, 15], and also using an ab
initio model [5]. In particular, the WSME model [12, 13]
uses two terms to construct an effective Hamiltonian and
partition function: (1) the free energy term associated
with the entropic cost of forming a pair of native residue
conformations with restricted dihedral angles and (2) an
enthalpic term associated with solvent-mediated contact
energies between residues. Thus, residues may be either
native or denatured, but not specific enough to distin-
guish sheets from helices. Our approach to polypeptides
is based on a Potts model, where residues could assume
many conformations including sheet, helix, coil, and turn.
Before discussing the full helix-coil-sheet system, let us
consider the simpler case of helix-coil transitions where
an effective (q = 2) Potts Hamiltonian (free energy in re-
ality) can be written for a protein consisting of N residues
as
− βHhc = h1
N
X
n=1
δ(1, in) + βJ1
N −1
X
k−1
Y
n=k+1
j=0
δ(1, in−j) (2)
where we assign in = 1 to a residue in helix confor-
mation and in = 2 to a residue in coil conformation.
The subscript 'hc' in −βHhc means 'helix-coil' and '1'
in h1 and J1 refers to helix. The meanings of these pa-
rameters are similar to those described in the WSME
model, where h1 < 0 refers to an entropic cost from
converting a coil to a helical residue, and J1 > 0 refers
to a contact energy between residues.
In the present
article, contact energies Ji are free-energies associated
with solvent-mediated interactions, including hydrogen
bonds, van der Waals, polar interactions, etc. The Kro-
necker delta δ(1, in) yields one if the nth residue is helical,
and zero otherwise. In the second term of Eq. (2), the
range k determines the range of interaction. In α-helices,
where k equals 3, residues at positions n − 3, n − 2, and
n − 1 are all helical when an H-bond forms between the
(n − 4)th and nth residues. Additionally, the (n − 4)th
and nth residues are not required to have the same con-
formation; in fact they could be in any conformation.
When k = 1, the effective Hamiltonian becomes −βHhc
= h1 PN
n=2 δ(1, in)δ(in−1, in). The
second term in Eq. (2) is also similar to the Hamilto-
nian of the GMPC model, which is a microscopic the-
ory for helix-coil transitions based on a q-state Potts
model [14, 16].
n=1 δ(1, in) + βJ1 PN
To write down an effective Hamiltonian suitable for β-
sheets, we need to include in it interactions up to length
L along the polypeptide chain. Such a Hamiltonian can
be constructed by adapting the long-range spin model of
Glumac and Uzelac [11]. For a chain of N spins, their
Hamiltonian can be written as
− βH =
L
N
X
X
l=1
n=1
βKlδ(in, in+l)
(3)
is distance-dependent. Fig. 1(a) illustrates
where Kl
a graphical representation of the L = 3 case and fa-
cilitates the construction of transfer matrices for long-
range Potts systems. For Potts systems on a 1D lat-
tice, Glumac and Uzelac grouped the spins along a chain
into columns of height L, the longest interaction length,
transforming a long-range problem of spin interactions
into a short range one relating nearest-neighbor columns
of height L [11, 17], illustrated in Fig. 1(b). Each column
of spins represents a vector that can take on one of qL
possible states. The transfer matrix thus has dimension
qL × qL. The various lines in Fig. 1 represent interac-
tions K1, K2, . . . , KL in Eq. (3), and contribute to the
(a)
i3
i2
i1
i6
i5
i4
2
j3
j2
j1
(b)
i9
i8
i7
i3
i2
i1
T
FIG. 1: (Color online) a) Graphical representation of the par-
tition function for the case L=3. The black dots mark the lo-
cations of particles along the chain. The dotted (blue) lines,
K1, are nearest neighbor interactions. The dashed (green)
lines, K2, are next nearest-neighbor interactions. Solid (red)
lines, K3, are the L=3 interactions. b) Graphical representa-
tion of the transfer matrix T .
partition function when the arguments in the Kronecker
delta's are equal.
Two modifications are made to apply the Glumac-
Uzelac method of constructing transfer matrices to a
protein system. Fig. 2(a) illustrates a segment of an
anti-parallel β-sheet, where interactions can occur be-
tween residues which are remote in relative chain posi-
tion, but are nearby in space. This is what is meant
by 'long-range'
in protein systems. Thus, the long-
range nature of a protein system comes from labeling
the residues according to the sequence order and does
not come from the spatial distance between two residues.
Even with the difference in the definition of long-range-
ness, the Glumac-Uzelac method can be used in solv-
ing the protein problem. The strengths of interactions
between each residue-residue pair are similar and not
dependent on the relative chain position l. This is a
main difference between our Hamiltonian (see Eq. (5) be-
low) and Eq. (3). For simplicity, in this article we shall
consider all contacts between β-strands are of the same
strength. In making this modification, Eq. (3) is recast
as −βH = βK PL
n=1 δ(in, in+l), which drops the
l-dependence of K, but maintains the long-range nature
of the Kronecker interactions.
l=1 PN
Secondly, according to Fig. 2(a), two hydrogen bonds
form between residue-residue pairs, which occur for ev-
ery other residue along a strand terminating at the turn.
On the other hand, the residues along the β-strand that
are not involved in hydrogen bonds with the opposite
β-strand, could be involved in hydrophobic interactions
with the opposite strand. To simplify the model, we as-
sume that every residue-residue pair along neighboring
strands forms contacts of the same strength, as stated
above. The following pattern then represents H-bonding
or hydrophobic interactions between two residues along
neighboring strands, which we identify as contacts: i1 →
i1+L, i2→i1+L−1,
In the
present work, the turn conformation is also counted as
a sheet conformation, but, in principle, the model can
be extended to include specifically turn conformations if
, i(L+1)/2→i(L+1)/2+1.
· · ·
3
(a)
i12
O
H
N
i11
O
H
N
i9
N
H
O
O
H
N
i10
N
H
O
i3
O
H
N
N
H
O
i8
i5
O
H
N
i2
N
H
O
i4
N
H
O
i1
H
N
i7
i6
O
O
N
H
(b)
i12
i9
i8
i5
i4
i1
(c)
i3
i6
i9
i12
i2
i5
i8
i11
i1
i4
i7
i10
U
V
U
i11
i10
i7
i6
i3
i2
FIG. 2: (Color online) a) A segment of an L = 11 anti-parallel β-sheet chain. The sequence position of a residue is labeled and
H-bonds are referenced by the dashed (red) lines. b) A simple pattern illustrating repeating L = 3 and nearest-neighbor contact
interactions, denoted by dashed (red) lines. The solid (black) lines represent peptide bonds. In (c), a diagram representing the
partition function for the structure in (b). The first column in (c) are residues i1, i2, i3, the second column are residues i4, i5, i6,
etc. Contacts are represented by dashed lines. The color of residues comprising the columns alternate in color from white to
black, which corresponds to the residue pattern in (b). Repeated multiplication of matrices U and V generates the partition
function for the whole chain.
q > 3. The Kronecker delta's given in Eq. (3) are then
modified to represent the aforementioned sheet-pattern.
Additionally, for protein systems where the neighboring
strands have the same interaction length L, the number
of strands M , and the total number of residues N are
related by
N = M R, R = (L + 1) /2
(4)
We can write the two-state effective Hamiltonian for a
pattern such as the one in Fig. 2(a) extended for any L,
while taking into account the two modifications made to
Eq. (3), as
− βHsc = h3
N
X
n=1
δ(3, in)
(5)
+ βJ3
R
M −1
X
X
k=1
m=1
b(ik,m)δ (cid:0)ik+R(m−1), i1−k+R(m+1)(cid:1) ,
where we denote in = 2 (coil), or 3 (sheet), b(ik,m) ≡
δ (cid:0)3, i1−k+R(m+1)(cid:1) and only allows J3 terms to accumu-
late when the residues at position k + R(m − 1) and
1 − k + R(m + 1) are locked in a sheet conformation and
are in contact. The term J3 > 0 now represents contacts
between sheet residues, h3 < 0 is the reduced entropic
cost for coil to sheet conversions. The subscript 'sc' in
−βHsc refers to 'sheet-coil' and subscript '3' in h3 and
J3 refers to sheet. Unlike in Eq. (2), we do not require
all residues between two residues in contact to be locked
into the sheet state.
To see the general pattern described by the second
term in Eq. (5), we start by considering the simplest
L = 3 case as shown in Fig. 2(b). In reality, the min-
imal structure in Fig. 2(b) may not even be considered
as a sheet structure, but nevertheless illustrates the gen-
eral behavior that the transfer matrix can be decomposed
into a product of sub-transfer matrices. For L = 3 case,
the transfer matrix decomposes into a product of two
matrices U and V , as illustrated by Fig. 2(b) and (c).
U and V are required to write out a general sequence
of M strands and are explicitly written with the help of
Fig. 2(c) as
hi U ji = xδ(i1,j1)+δ(i3,j3)+δ(j1,j2)
hi V ji = xδ(i2,j2)+δ(i3,j1)+δ(j2,j3)
(6)
where ii and ji are neighboring column vectors of length
L, where, for example, in Fig. 2(c), they can be hi =
hi1i2i3 and ji = i4i5i6i, and x = exp{βJ3}. Each
transfer matrix U and V has dimension qL × qL. This
methodology works for any finite L, where the total num-
ber of transfer matrices needed to generate a periodic
pattern for general L is found to be equal to the total
number of interactions over the distance L + 1, which
happens to equal the number R [18]. For example, for
the L = 3 case illustrated in Fig. 2(c), there are two in-
teractions, a nearest-neighbor (for example, i2, i3, in Fig.
2(c)) and one over the longest range of interaction (for
example, i1, i4, in Fig. 2(c)) thus two matrices are suf-
ficient. For illustrating purposes, we explicitly consider
a simple model of anti-parallel sheet-helix-coil systems,
which starts with a three-state (q = 3) effective Hamilto-
nian with four parameters that can describe transitions
between sheet, helix, and coil structures. Helical confor-
mations are assumed to form contacts between nearest
neighbors only, that is, the k = 1 case of Eq. (2). The
total effective Hamiltonian can be written as
− βHhcs = −βHhc − βHsc
(7)
where now in = 1, 2, or 3, refers to helix, coil, and sheet,
respectively, and the subscript 'hcs' in −βHhcs refers to
'helix-coil-sheet'. The partition function can be written
in the form of Eq. (1), when periodic boundary conditions
are imposed, and calculated using transfer matrices, sim-
ilar to the L = 3 case as illustrated in Fig. 2(b) and (c).
4
FIG. 3: (Color online) All calculated quantities using J1 = 2.85 kcal/mol, J3 = 2.45 kcal/mol, h1 = -4.91, and h3 = -4.20.
a) Order parameters for the case L = 11 with M = 100. b) Heat capacity (kcal/mol/K) vs. T for various strand lengths L
with M = 100. c) The same plot as in (b) with more details of the helix-sheet transition given. Black dots denote transition
temperatures which increase with range parameter L.
The parameters hi and Ji are chosen so that the he-
lix state is the most stable conformation at the lowest
temperature in the interested temperature range. The
coil dominates at high temperatures, where contact en-
ergies become relatively weak compared to thermal fluc-
tuations. The sheet is thus an intermediary state [19].
For some proteins, the sheet is seen as the most stable
conformation at low temperature, where the helix confor-
mation becomes an intermediary state [20]. Our model
can accommodate this case as well as a variety of others
with proper choices of parameters.
For systems with fixed numbers of residues, the parti-
tion function facilitates calculation of numerous thermo-
dynamical quantities, such as the average energy, hEi,
the heat capacity, C, and the order parameters, Θi, which
are the average fractional content of ith state among q
conformations at a particular temperature. To calcu-
late the partition function, we choose a multi-stranded
β-barrel system, which serves as an example of a protein
system satisfying periodic boundary conditions. Insert-
ing Eq. (9) into Eq. (1) and differentiating, we have for
such a system
C =
∂hEi
∂T
=
∂
∂T (cid:18)kB T 2 ∂ ln ZN
∂T (cid:19) and Θi =
∂ ln (ZN )
∂Ji
,
(8)
respectively. In Fig. 3(a), the order parameters for he-
lix, coil, and sheet are presented for the case L = 11,
M = 100, and in Fig. 3(b) and (c), we plot the tempera-
ture dependence of the heat capacity for various L cases.
The heat capacity curve show two peaks: the sharp, low-
temperature peak signifies the helix-sheet transition, and
the broad, high-temperature peak signifies the sheet-coil
transition. These peak positions are approximately given
by the crossing points of Θi, shown in Fig. 3(a), between
the helix and sheet and between the sheet and coil curves.
In conclusion, we have shown that, for a simple pat-
tern associated with anti-parallel β-sheet structures, an
effective Hamiltonian using a minimal number of param-
eters and its corresponding partition function can be con-
structed to study its helix-coil-sheet transitions. The
partition function can be exactly computed by means
of transfer matrices, which are used to calculate thermo-
dynamical properties of the system, including the order
parameters for helices and sheets and the heat capacity,
which show that increasing strand length, L, plays a sta-
bilizing role in the protein.
We would like to acknowledge Zvonko Glumac and
Katarina Uzelac for stimulating discussions and sending
us unpublished results. JMY wants to thank Professor
Sheng H Lin and Dr. A. N. Morozov for discussion and
support at the early stage of this work. We thank the
Pittsburgh Supercomputing Center for computing sup-
port.
[1] B. H. Zimm and J. K. Bragg, J. Chem. Phys. 31, 526
(1959); S. Lifson and A. Roig, J. Chem. Phys. 34, 1963
(1961).
[2] H. A. Scheraga, J. A. Vila, and D. R. Ripoll, Biophysical
Chemistry, 101-102, 255 (2002). A. J. Doig, Biophysical
Chemistry, 101-102, 281 (2002).
[3] D. Poland and H. A. Scheraga, Theory of Helix-Coil
Transitions in Biopolymers (Academic Press, New York,
1970).
[5] A.V. Yakubovich, I.A. Solov'yov, A.V. Solov'yov, and W.
Greiner, Eur. Phys. J. D. 46, 227(2008).
[6] S. Tanaka and H. Scheraga, Macromolecules 9, 812
(1976); S. Tanaka and H. Scheraga, Macromolecules 10,
9 (1977); S. Tanaka and H. Scheraga, Macromolecules
10, 305 (1977).
[7] H. Wako, N. Saito, and H. A. Scheraga, J. Protein Chem.
2, 221 (1983).
[8] W. L. Mattice and H. A. Scheraga, Biopolymers 23, 1701
[4] M. Takano, K. Nagayama, and A. Suyama, J. Chem.
(1984).
Phys. 116, 2219 (2002).
[9] J. K. Sun and A. J. Doig, J. Phys. Chem. B. 104, 1826
(2000).
[10] L. Hong and J. Lei, Phys. Rev. E. 78, 051904 (2008); L.
Hong, J. Chem. Phys. 129(22), 225101 (2008).
[11] Z. Glumac and K. Uzelac, J. Phys. A. 21, L421 (1988);
Z. Glumac and K. Uzelac, J. Phys. A. 22, 4439 (1989).
[12] V. Munoz, P. A. Thompson, J. Hofrichter, and W. A.
Eaton, Nature (London) 390, 196 (1997).
[13] V. Munoz, E. R. Henry, J. Hofrichter, and W. A. Eaton,
Proc. Natl. Acad. Sci. U.S.A. 95, 5872 (1998).; V. Munoz
and W. A. Eaton, Proc. Natl. Acad. Sci. U. S. A. 96,
11311 (1999); P. Bruscolini and A. Pelizzola, Phys. Rev.
Lett., 88, 258101 (2002).
[14] N. S. Ananikyan, Sh. A. Hajryan, E. Sh. Mamasakhlisov,
5
and V. F. Morozov, Biopolymers, 30, 357 (1990).
[15] R. E. Goldstein, Phys. Lett. 104A, 285 (1984).
[16] A. V. Badasyan, A. Giacometti, Y. Sh. Mamasakhlisov,
V. F. Morozov, and A. S. Benight, Phys. Rev. E., 81,
021921 (2010).
[17] Z. Glumac and K. Uzelac, J. Phys. A. 26, 5267 (1993).
[18] J. Schreck and J.M. Yuan, To appear
[19] F. Ding, J. M. Borreguero, S. V. Buldyrev, H. E. Stanley,
and N. V. Dokholyan, Proteins 53, 220 (2003).
[20] M. Andrec, A. K. Felts, E. Gallicchio, and R. M. Levy,
Proc. Natl. Acad. Sci. USA 102, 6801 (2005).
|
1212.1194 | 2 | 1212 | 2013-05-23T19:49:58 | Modelling the ATP production in mitochondria | [
"physics.bio-ph",
"q-bio.MN"
] | We revisit here the mathematical model for ATP production in mitochondria introduced recently by Bertram, Pedersen, Luciani, and Sherman (BPLS) as a simplification of the more complete but intricate Magnus and Keizer's model. We identify some inaccuracies in the BPLS original approximations for two flux rates, namely the adenine nucleotide translocator rate $J_{\rm ANT}$ and the calcium uniporter rate $J_{\rm uni}$. We introduce new approximations for such flux rates and then analyze some of the dynamical properties of the model. We infer, from exhaustive numerical explorations, that the enhanced BPLS equations have a unique attractor fixed point for physiologically acceptable ranges of mitochondrial variables and respiration inputs, as one would indeed expect from homeostasis. We determine, in the stationary regime, the dependence of the mitochondrial variables on the respiration inputs, namely the cytosolic concentration of calcium ${\rm Ca}_{\rm c}$ and the substrate fructose 1,6-bisphosphate FBP. The same dynamical effects of calcium and FBP saturations reported for the original BPLS model are observed here. We find out, however, a novel non-stationary effect which could be, in principle, physiologically interesting: some response times of the model tend to increase considerably for high concentrations of calcium and/or FBP. In particular, the larger the concentrations of ${\rm Ca}_{\rm c}$ and/or FBP, the larger the necessary time to attain homeostasis. | physics.bio-ph | physics | Bulletin of Mathematical Biology manuscript No.
(will be inserted by the editor)
Modelling the ATP production in mitochondria
Alberto Saa · Kellen M. Siqueira
3
1
0
2
y
a
M
3
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
4
9
1
1
.
2
1
2
1
:
v
i
X
r
a
Received: date / Accepted: date
Abstract We revisit here the mathematical model for ATP production in
mitochondria introduced recently by Bertram, Pedersen, Luciani, and Sher-
man (BPLS) as a simplification of the more complete but intricate Magnus
and Keizer's model. We identify some inaccuracies in the BPLS original ap-
proximations for two flux rates, namely the adenine nucleotide translocator
rate JANT and the calcium uniporter rate Juni. We introduce new approxi-
mations for such flux rates and then analyze some of the dynamical proper-
ties of the model. We infer, from exhaustive numerical explorations, that the
enhanced BPLS equations have a unique attractor fixed point for physiolog-
ically acceptable ranges of mitochondrial variables and respiration inputs, as
one would indeed expect from homeostasis. We determine, in the stationary
regime, the dependence of the mitochondrial variables on the respiration in-
puts, namely the cytosolic concentration of calcium Cac and the substrate
fructose 1,6-bisphosphate FBP. The same dynamical effects of calcium and
FBP saturations reported for the original BPLS model are observed here. We
find out, however, a novel non-stationary effect which could be, in principle,
physiologically interesting: some response times of the model tend to increase
considerably for high concentrations of calcium and/or FBP. In particular, the
larger the concentrations of Cac and/or FBP, the larger the necessary time to
attain homeostasis.
Keywords Mitochondria · Calcium · ATP · Mathematical model
Alberto Saa
Departamento de Matem´atica Aplicada,
Universidade Estadual de Campinas,
13083-859 Campinas, SP, Brazil
E-mail: [email protected]
Kellen M. Siqueira
Instituto de F´ısica "Gleb Wataghin",
Universidade Estadual de Campinas,
13083-859 Campinas, SP, Brazil
E-mail: [email protected]
2
Alberto Saa, Kellen M. Siqueira
1 Introduction
The exchange of energy in cells is mostly mediated by ATP (adenosine triphos-
phate) molecules. Such molecules are produced in several processes in an eu-
karyotic cell, but the principal source of ATP is typically the oxidative phos-
phorylation process which takes place in mitochondria. The mitochondrion
is an organelle with two membranes, having, therefore, two distinct bulk re-
gions: the intermembrane space and the mitochondrial matrix. In the inner
membrane, there are plenty of protein transporters and ionic channels, some
of which execute an active transport leading to a gradient of some ions and
molecules [1,2]. The metabolic cascade that leads to the production of ATP
in the mitochondrion starts in the cytoplasm. At first, glucose is transported
from the extracellular medium into the cytoplasm by GLUT transporters. It is
then converted in glucose-6-phosphate (G6P) by the enzyme hexokinase. G6P
is then converted in pyruvate in a process called glycolysis, in which there is a
net production of two ATP molecules. The pyruvate produced is transported
into the mitochondrion (to the mitochondrial matrix) and is metabolized in
a series of oxidation-reduction reactions in the citric acid cycle leading to the
production of the nicotinamide adenine dinucleotide NAD and flavin adenine
dinucleotide FAD. These electron donor molecules are oxidized in the com-
plexes I to IV present in the inner mitochondrial membrane. These reactions
lead to the activation of a proton pump, creating a pH gradient between the
inter membrane space and the matrix. The protons pumped into the inter-
membrane space return to the matrix through a transporter that uses their
energy to catalyze the conversion of ADP (adenosine diphosphate) into ATP.
The ATP produced in the mitochondria is then transported to the cytoplasm
by the ATP/ADP exchanger [1,2].
The kinetic aspects of the processes involved in the ATP production in
mitochondria are rather intricate. This issue was addressed by Magnus and
Keizer (MK), who introduced in the series of papers [3,4,5] a theoretical kinetic
model for ATP production in mitochondria based on the known biophysical
properties of the enzymes and transporters involved in the process. In fact, the
MK model was built by considering electrical activity and cytosolic calcium
handling in insulin-secreting pancreatic β-cells. The model consists basically
in a set of equations describing the dynamics of the citric acid cycle, the
proton pump, and the inner mitochondrial membrane transporters of ATP and
calcium. The MK model is effectively based on first biophysical principles and
provides a very detailed and accurate description of the processes considered to
be important for mitochondrial oxidative phosphorylation. However, it is also
a rather complex model with cumbersome equations, preventing a systematic
mathematical study of its main dynamical and physiological properties.
A simplification of the MK model aiming to retain its main dynamical prop-
erties was introduced recently by Bertram, Pedersen, Luciani, and Sherman
(BPLS) in [6]. The BPLS model incorporates some refinements introduced by
Cortassa et al. in [7] for the description of the ATP production in cardiac cells.
In fact, BPLS model can be considered as an approximation of the Cortassa et
Modelling the ATP production in mitochondria
3
al.'s model instead of the original MK one. As we will see, this was probably
the origin of some inaccuracies in the BPLS equations. As in the original MK
model, the mitochondrial ATP production in the BPLS model is governed by
four dynamical variables, namely the potential drop in the inner membrane
∆V and the mitochondrial concentrations of: reduced nicotinamide adenine
dinucleotide NADH, adenosine diphosphate ADP, and calcium Cam. The mi-
tochondrial concentrations of pyridine and adenine nucleotides are assumed to
be conserved
NADm + NADHm = NADtot,
ADPm + ATPm = Atot,
(1)
(2)
where NADtot and Atot stand for the total mitochondrial concentration of the
respective nucleotides. The balance of the pertinent fluxes and reactions yields
to the following dynamical equations for the mitochondrial variables
d
dt
NADH = JPDH − J0,
d
ADP = JANT − JF 1F 0,
dt
d
Cam = fm (Juni − JNaCa) ,
dt
d
∆V = C −1
dt
m (JH − JANT − JNaCa − 2Juni) ,
where
JH = JH,res − JH,ATP − JH,leak.
(3)
(4)
(5)
(6)
(7)
The derivation and meaning of the fluxes presented in the right-handed sides of
Eq. (3)-(6) are rather involved. The main details and the pertinent references
can be found, for instance, in the BPLS paper [6]. We have checked carefully
the derivation of each of these fluxes and we have found out some inaccuracies
in the BPLS expressions for the adenine nucleotide translocator rate JANT
and for the calcium uniporter rate Juni. As we will see, some of these problems
probably have originated in the transcription of the original MK equations to
the Cortassa et al.'s model.
In the present paper, we propose some enhanced approximations in the
BPLS framework for the fluxes JANT and Juni and analyze some of the dynam-
ical properties of Eqs. (3)-(6). We show, in particular, that for physiologically
acceptable ranges of mitochondrial respiration inputs, namely the cytosolic
concentration of calcium Cac and the substrate fructose 1,6-bisphosphate FBP,
the BPLS equations have a unique physiologically acceptable attractor fixed
point, as one would indeed expect for any model compatible with homeosta-
sis. Exhaustive numerical explorations indicate that the BPLS model is indeed
globally stable, reinforcing its relevance to physiological quantitative studies,
despite its simplicity when compared to the MK original model. We deter-
mine, in the stationary regime, the dependence on constant respiration inputs
Cac and FBP of the four mitochondrial variables considered in the model. As
4
Alberto Saa, Kellen M. Siqueira
in the original BPLS model, we observe here qualitatively distinct dynami-
cal behavior for low and high concentrations of Cac and/or FBP. We detect,
moreover, a non-stationary effect which could be, in principle, physiologically
interesting: the inertia of the system tends to increase considerably for high
concentrations of cytosolic calcium and FBP, i.e., some response times of the
model tend to increase considerably for high respiration inputs Cac and FBP.
In particular, the larger the concentrations of Cac and/or FBP, the larger the
necessary time to attain homeostasis.
2 The Enhanced BPLS Model
We will focus here in the problems we found for the BPLS expressions for the
adenine nucleotide translocator rate JANT and for calcium uniporter rate Juni,
since all the other quantities appearing in (3)-(6) were checked to be correct
and accurate for physiological ranges of variables and parameters. The MK
expression for the former is (see Eq. (16) and Table 4 of [3])
JANT = Vmax,ANT
1 − αc
αm
ATPc
ADPc
ATPc
ADPc
e−f F ∆V
ADPm
ATPm
F ∆V
RT
e−
RT (cid:17)(cid:16)1 + α−1
m
ADPm
ATPm(cid:17)
(cid:16)1 + αc
.
(8)
The precise meaning of all the quantities presented in this formula can be found
in [3,4,5], and in the BPLS paper [6] as well. (For the values of the parameters,
see Table 1). On the other hand, the expression for JANT presented in the Eq.
(35) of the Cortassa et al. paper [7] reads
JANT = Vmax,ANT
1 − αc
αm
ATPc
ADPc
ADPm
ATPm
(cid:16)1 + αc
ATPc
ADPc
e−f F ∆V
RT (cid:17)(cid:16)1 + α−1
m
ADPm
ATPm(cid:17)
.
(9)
By comparing with (8), we see clearly that it lacks the exponential in the
numerator. Furthermore, the incorrect expression (9) is transcribed in the
BPLS Eq. (35) as
JANT = Vmax,ANT
,
(10)
αm
ATPc
ATPm
ADPm − αc
ADPc(cid:17)(cid:16) ATPm
ADPm
ATPc
ADPc
+ α−1
m (cid:17) e−f F ∆V
RT
(cid:16)1 + αc
i.e., with another mistake in the denominator. The BPLS expression for JANT,
obtained from (10) after some simplifications, is
JANT = p19 ATPm
ATPm
ADPm
+ p20! ef F ∆V
ADPm
RT ,
(11)
where p19 and p20 are some (fitted) numerical parameters. The (reasonable)
physiological hypothesis used to derive (11) in the BPLS model is the assump-
tion that, due to the ion transporters action, the rates of ATP to ADP in the
Modelling the ATP production in mitochondria
mitochondrial matrix and in the cytoplasm are approximately the same,
ATPc
ADPc ≈
ATPm
ADPm
.
5
(12)
Note that this assumption implies from (8) that JANT ≈ 0 for ATPm →
Atot (and, hence, ADPm → 0 according to (2)), which is incompatible with
the BPLS expression (11). Another qualitatively different behavior arises for
larges values of ∆V : equation (8) implies that JANT tends to an asymptote,
whereas (10) suggests an exponential growth. The expression (10) is clearly
not accurate as an approximation of (8).
With the assumption (12) and taking into account the conservation of
mitochondrial pyridine nucleotides (2), the original MK expression (8) for the
adenine nucleotide translocator rate reads
JANT = Vmax,ANT(cid:18)
ATPm
Atot − (1 − αm)ATPm(cid:19)
αm − αce−
ATPm
Atot−ATPm
1 + αc
F ∆V
RT
. (13)
e−f F ∆V
RT
This is our first proposed approximation, which captures all the essential prop-
erties of (8) and is still simple enough to be mathematically manipulated.
Notice that for the typical range of physiological parameters, neglecting the
exponential in the numerator of (13) would imply a relative error inferior
to 5%. We will not, however, adopt this further approximation in this work.
Figure (1) illustrate the discrepancies between the expressions (11) and (13)
for typical physiological values of the parameters and variables. A closer in-
spection of the graphics (9) of [6] reveals that they have probably compared
their approximated expression (11) with Eq. (10), which was itself transcribed
incorrectly from Cortassa et al.'s Eq; (9).
With respect to the calcium uniporter rate Juni, the original MK expression
reads (see Eq. (19) in [3])
Juni = Vmax,uni
2F
RT (∆V − ∆V0)
1 − e−
RT (∆V −∆V0)
2F
Cac
Ktrans(cid:17)3
Ktrans (cid:16)1 + Cac
Ktrans(cid:17)4
(cid:16)1 + Cac
+
L
(1+Cac/Kact)na
. (14)
In the BPLS derivation of the approximation for Juni, it is used Eq. (38) of
Cortassa et al. [7], which reads
Juni = Vmax,uni
Cac
Ktrans (cid:16)1 + Cac
Ktrans(cid:17)4
(cid:16)1 + Cac
Ktrans(cid:17)3
(1+Cac/Kact)na (cid:16)1 − e−
+
2F
L
RT (∆V − ∆V0)
2F
RT (∆V −∆V0)(cid:17)
,
(15)
where one can see that there is a mistake in the denominator. The BPLS
proposed expression for the calcium uniporter rate, obtained as a simplification
of (15), is
Juni = (p21∆V − p22)Cac
2,
(16)
6
Alberto Saa, Kellen M. Siqueira
0.024
0.0235
0.023
0.0225
0.022
60
80
100
120
140
160
180
200
Fig. 1 Comparison between the original BPLS expressions and our proposals based on
the original MK model. Above: The BPLS adenine nucleotide translocator rate JANT (11)
and our proposal (13). We notice that the variation of (13) over physiological ranges (the
inserted graphics) is considerable smaller than that one of (11). Furthermore, even the
concavities of the curves are different. The dependency of (11) on ∆V is exponential, whereas
(13) tends to an asymptote for large values of ∆V . In accordance to the Table 1, these
curves were calculated by assuming ATP = 500 µM and Atot = 15 mM. Below: The BPLS
calcium uniporter rate Juni (16) and the original MK expression (14), both calculated for
Cac = 0.2 µM. Eq. (16) is a straight line which implies non-positive rates for physiological
values of ∆V .
where p21 ≈ 0.01 µM−1ms−1mV−1 and p22 ≈ 1.1 µM−1ms−1 are also fitted
numerical parameters. We found this equation to be inaccurate for the typical
physiological range of parameters as well (see Figure (1)). Notice, in particu-
lar, that it implies in non-positive flux rates for ∆V ≤ p22/p21 ≈ 110 mV. We
propose to keep in the approximated model the complete original MK equa-
tion (14). Its dependence on ∆V is already in a rather simple form, and the
complications for Cac are harmless for the dynamical studies, as we will show.
Modelling the ATP production in mitochondria
7
For the dynamical analysis, it is more conveniently to introduce the fol-
lowing dimensionless variables
x =
y =
z =
w =
u =
v =
,
NADHm
NADtot
ATPm
Atot
Cam
Ca0
∆V
∆V0
Cac
Ca0
FBP
FBP0
,
,
,
,
,
(17)
(18)
(19)
(20)
(21)
(22)
Taking into account the new proposed expressions (13) and (14), the rates in
the right-handed sides of (3)-(6) will be given by
JPDH = r1√v
x
J0 = r2
z
a1 + z (cid:18)a2 +
x
1 − x(cid:19)−1
,
(1 + a4ea5w)−1 ,
a3 + x
y
y
1 + a8
1 − a6y(cid:19) a7 − a8e−a9w
JANT = r3(cid:18)
1−y e−a10w ,
JF 1F 0 = r4(cid:2)(a11 + y)(cid:0)1 + a12e−a13w(cid:1)(cid:3)−1
JH,ATP = a15JF 1F 0,
JH,leak = r5(w − a16),
JNaCa = r6
JH,res = a14J0,
z
u
ea17w,
a18(w − 1)
1 − e−a18(w−1) G(u),
Juni = r7
where
G(u) =
u(1 + a19u)na(1 + a20u)3
a21 + (1 + a19u)na (1 + a20u)4 .
,
(23)
(24)
(25)
(26)
(27)
(28)
(29)
(30)
(31)
(32)
All the values of the numerical parameters and constants are presented in
Table 1. With the new dimensionless variables, the Eqs. (3)-(6) can be cast in
the form
x =
y =
1
NADtot
1
Atot
(JPDH − J0) ,
(JF 1F 0 − JANT) ,
(33)
(34)
8
Alberto Saa, Kellen M. Siqueira
NADtot = 10 × 103 µM
∆V0 = 91 mV
Vmax,ANT = 5 µM ms−1
f = 0.5
L = 110
αc = 0.111
r1 = 0.2 µM ms−1
r4 = 23.3 µM ms−1
r7 = 0.11 µM ms−1
a1 = 0.05
a4 = 4.23 × 10−16
a7 = 0.139
a10 = 1.68
a13 = 10.7
a16 = 0.16
a19 = 0.52
Atot = 15 × 103 µM
FBP0 = 1 µM
Vmax,uni = 10 µM ms−1
Ktrans = 19 µM
fm = 0.01
αm = 0.139
r2 = 0.6 µM ms−1
r5 = 0.182 µM ms−1
Ca0 = 0.2 µM
Cm = 1.8 µM mV−1
F
RT = 0.037 mV−1
Kact = 0.38 µM
na = 2.8
r3 = 5 µM ms−1
r6 = 0.001 µM ms−1
a2 = 1
a5 = 18.2
a8 = 0.111
a11 = 0.67
a14 = 11.7
a17 = 1.46
a20 = 0.01
a3 = 0.01
a6 = 0.861
a9 = 3.37
a12 = 5.10 × 109
a15 = 3.43
a18 = 6.73
a21 = 110
Table 1 Numerical parameters and rates for the enhanced BPLS model, see equations
(17)-(36). All the values were obtained from [3, 4, 5] and [6].
fm
Ca0
1
z =
w =
(Juni − JNaCa) ,
Cm∆V0
(JH − JANT − JNaCa − 2Juni) ,
(35)
(36)
where fm and Cm stand, respectively, for the fraction of free Ca ions and
the mitochondrial capacitance, see Table 1. Equations (33)-(36) form a non-
autonomous systems of four first order differential equations. The external
excitations u(t) and v(t) are related, respectively, to the cytosolic concentration
of calcium Cac and the substrate fructose 1,6-bisphosphate FBP, see Eqs. (21)
and (22). We can now start the dynamical analysis of the model.
3 Dynamics of the model
Let us consider initially the fixed points (x∗, y∗, z∗, w∗) of the system (33)-(36)
assuming constant inputs (u∗, v∗). By construction, the physiologically mean-
ingful range for the variables x and y is [0, 1], see (1)-(2) and (17)-(18). For z
and w, we assume only that they are non negative. The typical physiological
range for the potential drop, however, is more restrictive, corresponding to
∆V ≈ [90, 225] mV, which is equivalent to w ≈ [1, 2.5]. For the inputs u and
v, we consider the ranges [0, 10] and [0, 20], respectively, which corresponds to
Cac ≈ [0, 2] µM and FBP ≈ [0, 20] µM. We perform an exhaustive numerical
search [8] for fixed points of (33)-(36) by assuming u ∈ [0, 10] and v ∈ [0, 20]
constants. For all tested values of u and v, only one physiological (x, y ∈ [0, 1]
both z, w > 0) fixed point was found, which is always stable. Moreover, the
fixed point is globally stable for physiological ranges of variables, meaning that
any solution of (33)-(36) with reasonable initial conditions will tend asymptoti-
cally to the fixed point, i.e., the system indeed exhibits an asymptotic behavior
Modelling the ATP production in mitochondria
9
20
15
10
5
1
0
0
4
3
2
20
15
0.4
0.3
0.2
0.1
0
300
250
200
150
100
50
0
20
15
10
5
1
0
0
4
3
2
20
15
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
2
1.5
1
0.5
0
10
5
1
0
0
4
3
2
10
5
1
0
0
4
3
2
Fig. 2 The globally stable physiological fixed point (x∗, y∗, z∗, w∗) of the system (33)-(36)
assuming constant inputs (u∗, v∗).
compatible with homeostasis. Starting at a random point in the phase space,
the variables w and x have typically the quickest convergence to the fixed
point, where y and z are the slowest ones. The values of (x∗, y∗, z∗, w∗) as
function of the constant inputs (u∗, v∗) are depicted in Fig. (2), from where
one can already observe some physiologically consistent dynamical properties
which we describe in detail below.
The first observations is that the production of ATP and the concentration
of NADH vanishes in the absence of cytosolic calcium Cac and/or the substrate
fructose 1,6-bisphosphate FBP, i.e., x∗ and y∗ → 0 for u∗ or v∗ → 0. Notice
that, from the condition JNaCa = Juni defining the fixed point z∗ = 0 (see Eq.
(35)), we have z∗ = 0 for u∗ = 0. On the other hand, JPDH vanishes for z∗ = 0
(and for v∗ = 0 as well), which implies via the condition x∗ = 0 that J0 = 0
and, consequently, x∗ = 0. The condition for y∗ and w∗ are more involved.
The former vanishes for vanishing u∗ or v∗, while the latter will be given by
w∗ ≈ a16 = 0.16 for u∗ = 0. Also, we see that for reasonable values of u∗
and v∗ the value of the potential drop ∆V (w∗) is almost constant and close
to 150 mV (w∗ = 1.65). This stability is probably the reason why the original
BPLS model is robust, despite the inaccuracies for the expression of JANT and
Juni we are correcting in this paper. We will return to this point in the last
section. Still from the condition JNaCa = Juni, we see that z∗ ∝ u∗G(u∗), since
w∗ is almost constant for physiological reasonable values of u∗ and v∗ (see Fig.
(2c)).
10
Alberto Saa, Kellen M. Siqueira
Another important feature of the BPLS model is the reversion of the dy-
namical behavior of some mitochondrial variables in the presence of lower and
higher concentration of cytosolic calcium and FBP. This behavior can be seen,
for instance, in Fig. (2b). After attaining its maximum, the ATP production
(y∗) tends to decrease for increasing cytosolic calcium concentrations (u∗).
Calcium saturation can be simulated, as described in [6], by setting a1 = 0
in the expression for JPDH (23). The reversion of the dynamical behavior of
the other mitochondrial variables for higher Cac concentrations can also be
inferred directly from Fig. 2, but it is certainly better illustrated in Fig. 3,
which depicts the solutions of (33)-(36) for an oscillatory Cac input of the
form
u(t) = u0 + u1 sin(t/t0),
(37)
with constant v(t) and initial conditions (x(0), y(0), z(0), w(0)) given by the
values of the fixed point corresponding to u∗ = u(0) and v∗ = v(0). As we
will see, such a choice of initial condition is consistent with the adiabatic
(stationary) regime we observe for sufficiently slow inputs (large periods t0).
For lower values of u0 (low Cac concentrations), all the mitochondrial variables
increases and decreases in synchrony with the variations of u. On the other
hand, for higher values of u0, the dynamical behavior of x, y and w is reversed.
i.e., they tend to decrease/increase while u increases/decreases. This effect can
be understood from the relation between u∗ and z∗ depicted in Fig. (2c). The
value of z∗ tends to increase rapidly when u∗ increases and, for large values of
z, the dependence of the expression for JPDH (23) on z saturates and becomes
equivalent to setting a1 = 0. Without the z suppression term in JPDH, the
dynamical behavior of the variables x, y, and w is reversed, as it was pointed
out in the original BPLS analysis. Low variations of v (FBP) do not change
qualitatively this dynamical behavior. However, the situation changes for large
concentrations of FBP. As described in [6], for low concentrations of FBP, the
NADHm concentration reacts to a sudden rising of Cac with an upward teeth,
while for high concentrations of FBP such behavior is reversed, i.e., NADHm
concentration exhibits a downward teeth if Cac increases. This situation is
analyzed and depicted in Fig. (4).
The oscillatory excitations used in the examples depicted in Fig. (3) and
(4) have period t0 = 3 min. For inputs varying over a time scale of minutes,
the system evolves adiabatically in a good approximation, i.e., the instanta-
neous solution (x(t), y(t), z(t), w(t)) is well approximated by the fixed point
(x∗, y∗, z∗, w∗) corresponding to u∗ = u(t) and v∗ = v(t). In other words, for
slowly varying inputs, the solutions of the system are confined to the fixed-
point surfaces depicted in Fig. 2. Of course, one expects a breakdown of this
adiabatic behavior for rapidly varying inputs. Non-stationary effects must ap-
pear for inputs varying with a characteristic time smaller than a certain critical
value. In order to study non-stationary effects in our model, we consider the
response of the system for inputs of the type
t0 (cid:19) ,
u(t) = u0 + u1 tanh(cid:18) t − t1
(38)
Modelling the ATP production in mitochondria
11
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
0.82
0.815
0.81
0.805
0.8
0.795
0.79
0.785
0.78
30.74
30.738
30.736
30.734
30.732
30.73
30.728
2.05
2
1.95
1.9
1.85
1.8
1.75
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
1.7
0
1
2
3
5
4
time (min)
6
7
8
9
10
0.1
0.095
0.09
0.085
0.08
0.075
0.07
0.065
0.06
24
22
20
18
16
14
4
3.8
3.6
3.4
3.2
3
2.8
2.6
0.04
0.035
0.03
0.025
0.02
0.015
151
150
149
148
147
146
145
144
0.01
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
29
28.5
28
27.5
27
26.5
26
25.5
25
24.5
24
142.5
142
141.5
141
140.5
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
Fig. 3 Response of the equations (33)-(36) to oscillatory inputs (37). Notice that for low Cac
concentrations (left), all the mitochondrial variables increases and decreases in synchrony
with the variations of u. On the other hand, for high Cac concentrations (right), the behavior
of x, y and w is reversed. All the curves were evaluated for FBP = 0.5 µM. See the text for
further details.
for different values of t0. This situation is depicted in Fig. 5 for some values
of u0 and u1 and for t0 = 2.5, 1, 0.5, and 0.02 seconds. It is clear that for lower
values of u (Cac), approximately 10 seconds are enough to assure that NADHm
concentration and ∆V reaches their values corresponding to the adiabatic
regime, which in this case corresponds to the homeostasis. As we have already
noticed, the variables y (ATP) and z (Cam) are the slowest ones to attain
their respective stationary regimes. For lower values of f u (Cac), they spend
12
Alberto Saa, Kellen M. Siqueira
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0
1
2
3
5
4
time (min)
6
7
8
9
10
31
30
29
28
27
26
25
24
23
4.5
4
3.5
3
2.5
2
1.5
1
0.5
0
200
150
100
50
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
0
1
2
3
5
4
time (min)
6
7
8
9
10
160
150
140
130
120
110
100
90
80
0
1
2
3
5
4
time (min)
6
7
8
9
10
2 100
2 000
1 900
1 800
1 700
1 600
1 500
1 400
1 300
1 200
11
10
9
8
7
6
5
4
3
2
1
450
400
350
300
250
200
150
100
50
0
170
165
160
155
150
145
140
135
130
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
0
1
2
3
5
4
time (min)
6
7
8
9
10
Fig. 4 Response of the equations (33)-(36) to oscillatory inputs (37) for different concen-
trations of FBP. The system is submitted to the oscillatory input corresponding to the top
graphic. The left column is the response for FBP = 0.5 µM, while the right one corresponds
to FBP = 10 µM. Its clear that the NADHm concentration reverses its dynamical behavior
for low and high concentrations of FBP. See the text for further details.
approximately 40 seconds to stabilize. Increasing the values of u implies the
increasing of such "relaxation" times, i.e., a larger time is necessary to attain
homeostasis. The second column in Fig. 5 corresponds to a situation with
u ∈ [2, 4], for which almost 30 seconds are necessary to assure the attainment
of the stationary regime for the rapid variable w, whereas the slow one will
Modelling the ATP production in mitochondria
13
0.22
0.2
0.18
0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
30
25
20
15
10
4 500
4 000
3 500
3 000
2 500
2 000
1 500
1 000
0
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
0.85
0.8
0.75
0.7
0.65
0.6
0.55
0.5
0.45
0.4
0.35
30.82
30.8
30.78
30.76
30.74
30.72
30.7
30.68
4 000
3 500
3 000
2 500
0
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
0.3
0.25
0.2
0.15
0.1
0.05
0
0
25
20
15
10
5
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
154
152
150
148
146
144
142
140
138
136
134
0
5
10
15
20
time (sec)
25
30
35
40
152
150
148
146
144
142
140
0
5
10
15
20
time (sec)
25
30
35
40
Fig. 5 Response of the equations (33)-(36) to step-like inputs (38). The magenta, blue, red,
and green curves correspond, respectively, to t0 = 2.5, 1, 0.5, and 0.02 seconds. The inertia
of the system increases considerably for higher values of u, see the text for further details.
The curves were evaluated for FBP = 0.5 µM.
need a few minutes. The inertia of the system, hence, increases considerably
for higher concentrations of cytosolic calcium.
Higher concentrations of fructose 1,6-bisphosphate FBP also imply an in-
creasing of the inertia of the system. This situation is analyzed and depicted
in Fig. (6). Besides the increasing of the relaxation times, for higher concen-
trations of FBP we observe a smoothing out of the dynamical response of
NADHm. In particular, its overshooting present for large variations of CAc
14
Alberto Saa, Kellen M. Siqueira
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
30.8
30.78
30.76
30.74
30.72
30.7
30.68
30.66
30.64
4 500
4 000
3 500
3 000
2 500
2 000
1 500
1 000
5
10
15
20
time (sec)
25
30
35
40
5
10
15
20
time (sec)
25
30
35
40
2 200
2 100
2 000
1 900
1 800
1 700
1 600
1 500
1 400
1 300
11 000
10 000
9 000
8 000
7 000
6 000
5 000
4 000
3 000
0
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
250
200
150
100
50
0
0
165
160
155
150
5
10
15
20
time (sec)
25
30
35
40
500
0
120
100
80
60
40
20
0
0
160
150
140
130
120
110
100
90
0
5
10
15
20
time (sec)
25
30
35
40
0
5
10
15
20
time (sec)
25
30
35
40
Fig. 6 Response of the equations (33)-(36) to step-like inputs (38). The magenta, blue, red,
and green curves correspond, respectively, to t0 = 2.5, 1, 0.5, and 0.02 seconds. The inertia of
the system also increases considerably for higher values of v, see the text for further details.
The system is submitted to the input corresponding to the top graphics. The left column
depicts the response for FBP = 0.5 µM, while de right one corresponds to FBP = 10 µM.
See the text for further details.
and low FBP disappears for the high FBP concentration case, compare Fig-
ures (5) and (6). By examining this overshooting in the dynamics of NADHm,
our rapidest variable, it is possible to estimate the critical time for which any
adiabatic approximation should break. We can see from Figures (5) and (6)
that the overshooting appears for transitions occurring in less than 2.5 seconds
Modelling the ATP production in mitochondria
15
approximately. We do not expect any stationary response for input variables
varying over periods smaller than this.
4 Final Remarks
We have revisited here the mathematical model for ATP production in mi-
tochondria introduced recently by Bertram, Pedersen, Luciani, and Sherman
(BPLS) in [6] as a simplification of the more complete but intricate Magnus
and Keizer's model [3,4,5]. We checked carefully all the approximations intro-
duced in the BPLS model and found some inaccuracies for the approximations
used for the adenine nucleotide translocator rate JANT and for the calcium uni-
porter rate Juni. We proposed some enhanced approximations for such rates
based on the original Magnus and Keizer's model and analyzed some dynamical
properties of the model. Our results for the stationary regime indicate that the
BPLS model is indeed globally stable, reinforcing its relevance to physiologi-
cal quantitative studies, despite its simplicity when compared to the Magnus
and Keizer's model. We have considered also the non-stationary regime and
detected a effect which could be, in principle, physiologically interesting: the
inertia of the system tends to increase considerably for high concentrations
of cytosolic calcium and FBP, i.e., some response times of the model tend to
increase considerably for high respiration inputs Cac and FBP. In particular,
for Cac ≈ 0.2 µM and FBP ≈ 0.5 µM, approximately 10 seconds are necessary
to NADHm and ∆V attain homeostasis after a sudden increasing in Cac. The
variables ATPm and Cam are typically slower and need approximately 30 sec-
onds to attain homeostasis in the same conditions. Keeping FBP constant and
increasing Cac, or keeping Cac and increasing FBP, will imply a considerably
increasing of this response time, i.e., the system will take a longer time to
attain homeostasis.
It is interesting to notice that the dynamics of our enhanced model are
qualitatively similar to the original BPLS one, despite the differences in the
rates JANT and Juni for physiological ranges, as depicted, for instance, in
Fig. 1. This point can be understood from the fact that the value of w∗,
which does not depends tightly on the details of such rates, is almost constant
and corresponding to ∆V = 150 mV for reasonable values of the inputs v∗
and u∗. For a fixed value of ∆V , the numerical parameters in (11) can be
fitted to provide a good adjustment for the real ATP dependence of (13).
An inspection of Fig. 9 of [6] reveals that the adjustment of their numerical
parameters was probably checked for ATP ≈ 3 mM and for ∆V ≈ 160 mV,
which is close to the physiological global fixed point (homeostasis), explaining
why the asymptotic dynamics are not strongly affected by the inaccuracies
in the BPLS approximations. On the other hand, we do not expect that the
detected non-stationary effects be independent on the details of JANT and
Juni. Such points certainly deserve further investigations.
16
Alberto Saa, Kellen M. Siqueira
Acknowledgements The authors are grateful to FAPESP and CNPq for the financial
support. AS wishes to thank Prof. Leon Brenig for the warm hospitality at the Brussels
Free University, where part of this work was carried on.
References
1. A. C. Guyton and J. E. Hall, Textbook of Medical Physiology, Elsevier (2006).
2. D. L. Nelson and M. Cox, Lehninger Principles of Biochemistry, W. H. Freeman and
Company (2004).
3. G. Magnus and J. Keizer, Am. J. Physiol. 273, C717-C733 (1997).
4. G. Magnus and J. Keizer, Am. J. Physiol. 274, C1158-C1173 (1998).
5. G. Magnus and J. Keizer, Am. J. Physiol. 274, C1174-C1184 (1998).
6. R. Bertram, M.G. Pedersen, D.S. Luciani, and A. Shermand, J. Theor. Biol. 243, 575-
586 (2006).
7. S. Cortassa, M.A. Aon, E. Marban, R.L. Winslow, and B. O'Rourke, Biophys. J. 84,
2734-2755 (2003).
8. Scilab files are available at http://vigo.ime.unicamp.br/atp
|
1202.3037 | 2 | 1202 | 2012-03-11T12:47:25 | Landing together: how flocks arrive at a coherent action in time and space in the presence of perturbations | [
"physics.bio-ph"
] | Collective motion is abundant in nature, producing a vast amount of phenomena which have been studied in recent years, including the landing of flocks of birds. We investigate the collective decision making scenario where a flock of birds decides the optimal time of landing in the absence of a global leader. We introduce a simple phenomenological model in the spirit of the statistical mechanics-based self-propelled particles (SPP-s) approach to interpret this process. We expect that our model is applicable to a larger class of spatiotemporal decision making situations than just the landing of flocks (which process is used as a paradigmatic case). In the model birds are only influenced by observable variables, like position and velocity. Heterogeneity is introduced in the flock in terms of a depletion time after which a bird feels increasing bias to move towards the ground. Our model demonstrates a possible mechanism by which animals in a large group can arrive at an egalitarian decision about the time of switching from one activity to another in the absence of a leader. In particular, we show the existence of a paradoxical effect where noise enhances the coherence of the landing process. | physics.bio-ph | physics |
Landing together: how flocks arrive at a coherent action
in time and space in the presence of perturbations
B. Ferdinandya,∗, K. Bhattacharyaa,b, D. ´Abela, T. Vicsek a,c
aDepartment of Biological Physics, Eotvos University, Budapest, H-1117 Hungary
bDepartment of Physics, Birla Institute of Technology and Science, Pilani 333031,
cStatistical and Biological Physics Group, ELTE-HAS, P´azm´any P. Stny. 1A,
Budapest, H-1117 Hungary
India
Abstract
Collective motion is abundant in nature, producing a vast amount of phenomena
which have been studied in recent years, including the landing of flocks of birds.
We investigate the collective decision making scenario where a flock of birds de-
cides the optimal time of landing in the absence of a global leader. We introduce a
simple phenomenological model in the spirit of the statistical mechanics-based self-
propelled particles (SPP-s) approach to interpret this process. We expect that our
model is applicable to a larger class of spatiotemporal decision making situations
than just the landing of flocks (which process is used as a paradigmatic case). In the
model birds are only influenced by observable variables, like position and velocity.
Heterogeneity is introduced in the flock in terms of a depletion time after which a
bird feels increasing bias to move towards the ground. Our model demonstrates a
possible mechanism by which animals in a large group can arrive at an egalitarian
decision about the time of switching from one activity to another in the absence of
a leader. In particular, we show the existence of a paradoxical effect where noise
enhances the coherence of the landing process.
Key words:
Collective motion, Flocking, SPP model, Group decision making, Landing
PACS: 87.15Zg, 87.19lo
∗ Corresponding author. E-mail: [email protected]
Preprint submitted to Elsevier
11 November 2018
1
Introduction
Scenarios involving collective decision making appear to be ubiquitous in sev-
eral fields including animal behaviour [1] and social sciences [2]. Several stud-
ies [3,4] have investigated the cases where any group unanimously decides to
choose one of the many available options. It is generally expected that there
will be differences in the motivations of the members at the time when mak-
ing choices. However, in spite of the differences consensus is seen to evolve.
Different mechanisms have been suggested to account for this type of process
[5,6,7]. The related phenomenon of collective motion in animal groups has also
been extensively modelled [8,9,10] by considering animals as point-like parti-
cles (so-called self-propelled particles). In these models it is assumed that each
individual moves with constant speed while tending to align with its imme-
diate neighbours, for low level noise, giving rise to a globally ordered state
replicating the motion of flocks where animals move in the same direction.
Our object of study is the phenomenon during which the animals moving in
groups seemingly make unanimous decisions on the time and choice of per-
forming activities [1,11] even in the absence of global leaders. Examples include
takeoff of swarm of honeybees from nest sites [12], activity synchronisation in
Merino sheep [13], collective movement of white-faced capuchins [14], group
departures of domestic geese [15] and departure of Argentine ants from feed-
ing site [16]. Such a paradigm could be generalised to human behaviour as
well, e.g. where to stop to rest, when making an excursion with a group. The
important extra feature of taking into account group motion during collective
decision making is that in such an approach the neighbourhood with which a
consensus is to be achieved is dynamically changing in a realistic manner.
In this report we model the process of landing of bird flocks performing forag-
ing flights as a typical example of collective decision making. We regard the
birds as self-propelled particles, the only difference between the birds being an
a priori value corresponding to the heterogeneity in motivations. Throughout
this paper we shall use the terms "birds" and "particles" interchangeably. In
a recent model for collective landing [17] the birds are assumed to move under
the action of different social forces [18]. In addition, the internal state of each
bird is characterised by a continuous variable called landing intent such that
the internal state of each bird is directly coupled to the internal state of its
neighbours. Another model [19] allows the motivation of individual birds to be
influenced by only the observable variables of their neighbours, like velocity
and position. This assumption simplifies the description of the landing pro-
cess, since it does not presume very sophisticated channels of communication
during flight. The model discussed in this report has also been formulated
along similar lines. The decision of a bird is influenced by the state of mo-
tion of its neighbours and an a priori intent (a pre-assigned constant) taking
2
into account its endurance that acts through a noise variable. This makes the
landing a stochastic process even if we consider a single bird only.
Our aim with this report is to quantitatively show that a group without leaders
can perform a synchronised landing following very simple rules, even when a
single bird does not have information about the whole of the flock. One of the
important new features of our approach is the possibility of tuning the level
of perturbations that the particles are subject to during the landing process.
This approach also allowed us to uncover an interesting effect. According to
our simple model, there is an optimal level of perturbations resulting in the
smallest spread of the landing times of the individual birds, meaning that in
certain cases the noise can enhance the coherence of the landing process.
2 Model
2.1 Basics
In this section, we will describe the basic concepts of our model, leaving the
majority of the technical details to the next section. We formulate the dy-
namics of a bird flock along the lines of social forces. These forces are similar,
but not the equivalent of forces considered in earlier studies [8,20,21]. In the
model, we consider a flock that performs a horizontal flight about an average
height until it decides to land. This allows us to handle the horizontal and
vertical motion of the birds separately.
xy - a repulsive force, and (iv) f c
xy - a cohesive force. The force f a
Let us first regard the horizontal motion. Setting z = 0 as the only natural
boundary, representing the ground, the horizontal flight is performed parallel
to the xy-plane. There are four social forces acting upon the ith bird horizon-
tally, (i) f a
xy - an averaging on the velocities of other birds in its neighbourhood,
(ii) f n
xy - an evenly distributed noise i.e., a random perturbation on its veloc-
ity, (iii) f r
xy is
responsible for aligning the velocities of neighbouring birds in the flock. For
this purpose we define the neighbourhood (Ni) of the ith bird as those birds
that are encompassed by a cylinder of infinite height and radius R centred
on the ith bird. While in flight a bird would like to maintain some separation
with other birds. This is accounted by the repulsive force. This force is char-
acterised by a certain radius of interaction d, which can also be considered
as the effective "size of a bird" and is much lower than the interaction radius
characterising Ni. The purpose of the cohesive force is to keep the flock from
breaking apart. It is described by another infinite cylinder with a radius D
which is much larger than that of Ni, and is centred on the centre of mass
(CoM) of all N birds. If a bird strays out of this cylinder it is propelled towards
3
the centre of mass.
z - a noise. The force f h
The vertical motion also consists of four social forces: (i) f a
z - an averaging on
the vertical velocities of the neighbourhood Ni of the ith bird, (ii) f h
z - a force
that tries to keep the birds about a given height, (iii) f r
z - a repulsive force
like in the vertical direction, and (iv) f n
z is such, that
around a given height h there is a regime of width ∆h where in all practical
sense, a bird moves freely, but outside of that, the bird is repelled toward h.
The most important part of the model is the vertical noise, since this is the
force that facilitates the landing of the birds. This is done as follows: each
bird is assigned an a priori value, a ti depletion time, which represents the
time at which the bird starts to feel the depletion of its energy reserves, and
would increasingly wish to land. In general these depletion times will depend
on several external or internal conditions such as the energy reserves of the
birds [22], stamina, health, willingness to fly and other things and thus will be
in general different for different birds. As such, these values are drawn from a
Gaussian distribution of a mean µ and a variance σ2. When a bird i reaches its
ti, the originally evenly distributed f n
z noise starts to get increasingly biased
to facilitate landing with a characteristic time τ . Landing occurs when this
force and the averaging force overwhelms f h
z , allowing the birds to land. Thus
the equations of motion of the ith bird in the horizontal direction are:
f sum
xy,i = f a
xy,i + f n
xy,i + f c
xy,i + f r
xy,i
and
rxy,i(t + ∆t) = rxy,i(t) + v
f sum
xy,i
f sum
xy,i ∆t,
(1)
(2)
where rxy,i is the position of the ith bird in the xy-plane, and v is a constant,
so that the birds move with a velocity of constant magnitude. Very similar
equations hold for the vertical motion:
f sum
z,i = f a
z,i + f n
z,i + f h
z,i + f r
z,i
zi(t + ∆t) = zi(t) + v
f sum
z,i
f sum
z,i
∆t,
(3)
(4)
where zi is the vertical position of the bird. As seen in the next subsection, the
magnitude of these forces are parameters of the model, which seemingly are
a lot in number, but only some of them are actually relevant to the problem
of landing. Therefore we study the behaviour of the system as a function of a
few parameters, viz. the variance of depletion times σ, the number of birds N ,
and the coefficients of f h
z,i (described in 2.2). The characteristic time
τ is only important in the sense that it must be of the order of magnitude of σ
for non-trivial behaviour of the model. In addition, all the differences between
the birds are contained in σ.
z,i and f n
4
2.2 Details
In this section we provide the exact mathematical expressions for the forces,
and other details concerning the model. Noting that vxy and vz stand for
the instantaneous velocities in the horizontal plane and the vertical direction
respectively, the averaging forces are given by:
(6)
where (cid:104) (cid:105)Ni
is the average over the birds in the neighbourhood Ni of the ith
bird. We also assume that the averaging includes birds which have already
landed (in Ni) so that their influence is taken into account. The effect of the
landed birds on the averaging forces would be the same as that of those moving
vertically downwards. The repulsive forces follow the same scheme in all three
dimensions. The explicit form for the z-direction is as follows
xy,i = (cid:104)vxy(cid:105)Ni
f a
z,i = (cid:104)vz(cid:105)Ni
f a
N(cid:88)
j=1
f r
z,i =
f r
z,ij
(5)
(7)
(8)
(9)
where
f r
z,ij =
A (d − zi − zj) if 0 < zi − zj < d
if r∆CoM
otherwise,
(cid:16)r∆CoM
otherwise.
−B
≤ D
2
(cid:17)
xy,i
− D
2
xy,i
0
0
f c
xy,i =
Here A is the strength of the repulsive force. The cohesive force on the ith
bird is given by the following equation:
xy,i
where r∆CoM
is its distance from the CoM and B is the strength of the force.
The form of the force ensures that if a bird is farther from the CoM than D/2
it feels an attraction towards the CoM. The social force describing the will to
stay at a given height h is given by
(cid:34)
(cid:40)10
R
(cid:32)
zi − h − ∆h
2
(cid:33)(cid:41)(cid:35)
z,i = − C
f h
20
1 + tanh
sign(zi − h),
(10)
where C is the strength of the force. The attraction f h
z,i towards the preferred
altitude h is actually weak within a region of width ∆h and is strong outside
(see figure 1a).
We now define the nature of the noise in the vertical and the horizontal di-
rections. At any instant of time the ith bird is influenced by a horizontal
5
(a)
(b)
Fig. 1. (a) The force f h
z against z. The plot shows the small forceless regime around
h and the fast strengthening of the force outside of that, quickly saturating to a
constant. (b) The plot shows the time evolution of the probability distribution of
f n
z . Equation 13 is the explicit formula for generating values of this force.
noise:
(11)
where ξxy,i is a unit vector on the xy-plane whose orientation is taken to be
random and β is the strength of the noise. For times less than ti the vertical
noise fz,i is given by:
f n
xy,i(t) = βξxy,i(t),
(12)
where ξz,i(t) is randomly chosen from a uniform distribution on the interval
[−1, 1] and α is the amplitude of the vertical noise. After time t crosses ti the
form of f n
f n
z,i(t) = αξz,i(t),
1 + 4tanθ + 4tan2θ − 8ξz,i(t)tanθ
(cid:26)
(cid:19)(cid:27)
2tanθ
(cid:18)
−t − ti
τ
.
1
2
1 + exp
(13)
z,i gets modified to:
α − α
(cid:113)
f n
z,i(t) =
where
tanθ =
The above form ensures that after ti the noise becomes continually biased
towards the downward direction (see figure 1b). The characteristic time scale
for this biasing is given by τ .
For setting the units of measurement we choose R = 1 and ∆t = 1. The
summary of the parameters, and their typical values, as used in our simulations
are provided in table 1. As initial condition the birds are assigned randomly
oriented velocities and are uniformly distributed in a cylinder of diameter D
and of height ∆h.
For a rough fitting of the model to real flocks we took data from [23]. Based
upon their measurements it seems reasonable to take v ≈ 10 m/s and R ≈ 2 m,
6
Table 1
The parameters with their brief explanation and values as we most often used them
in our simulations. We find that the most relevant parameters to the problem of
landing are α, C, σ and N . The values of the parameters were chosen so to obtain
a biologically relevant landing scenario. For setting the units of measurement we
choose R = 1 and ∆t = 1, which makes the horizontal interaction radius (the
radius of N ) 1.
Param.
Description
Value Dimension
the coefficient of the vertical noise
the coefficient of f h
z
0.2095
0.81
the standard deviation of the distribution of ti
5000
R/∆t
R/∆t
∆t
-
R/∆t
R
R
R
R
1/∆t
∆t
1/∆t
1/∆t
300
0.1
20
1/6
100
2
5000
50000
100
100/3
α
C
σ
N
v
D
d
h
τ
µ
A
B
β
∆h
the effective width of the flock
the number of birds
the speed of the birds
the horizontal diameter of the flock
the "size" of the bird
the optimal height of flight
the timescale of energy depletion
the mean of the distribution of ti
the coefficient of f r
z and f r
xy
the coefficient of f c
xy
the coefficient of the horizontal noise
0.05
R/∆t
which makes ∆t ≈ 0.02 s, since v = 0.1 R/∆t in our model. As a rough
estimate from the simulations, we can take the time needed for the flock to land
as 5000 ∆t. This means that our simulated flocks flying at the altitude of 200 m
land in about a 100 s. Taking into account, that our model is very simplified
compared to the actual dynamics of a bird flock, the different parameters of
the model are in a reasonable range for real-life bird flocks.
2.3 Quantitative characterisation of the landing
To quantitatively characterise the cohesion of the landing, we introduce a few
quantities. We measure the spatial cohesion of the flock by taking the standard
deviation (σxy) of the horizontal coordinates from the centre of mass as the
7
following:
where
(cid:118)(cid:117)(cid:117)(cid:117)(cid:116) 1
N
σxy =
(cid:16)
N(cid:88)
rxy,j − rCoM
(cid:80)N
j=1
xy
rCoM
xy =
j=1 rxy,j
.
N
(cid:17)2
(14)
We sample this quantity at two different times, once during flight (σ0
xy) and
once after the whole flock has landed (σxy), so that their ratio forms a measure
to characterises the degree of coherence of the landing in space.
To characterise the temporal coherence of the landing we measured the fol-
lowing two quantities viz. the standard deviation of the times at which the
different birds have landed (σL) and the time (T60) that elapses between the
landing of the 20% of the flock and the landing of the 80% of the flock. We
additionally calculated the latter quantity in the case when the coupling be-
tween the birds was set to zero, i.e., when f a
60. To obtain
normalised measures characterising the temporal coherence we took σL/σ and
T60/T 0
60.
z = 0 denoted by T 0
2.4 Ordering horizontally
For obtaining a biologically relevant scenario, we adjusted the coefficient of
the horizontal noise β to an appropriate value. As can be seen in figure 2,
that starting from the totally coherent, no noise situation, the increase in β
gradually destroys this coherence leading to a situation, where the centre of
mass of the flock does not move at all. In this case we see the boundaries of
f c
xy if we plot the tracks of the birds projected unto the xy-plane. We chose
to investigate the regime where the horizontal tracks would be similar to the
one illustrated in figure 2b.
2.5 Assay of the landing
After choosing the parameters for horizontal motion we observe the time evo-
lution of the z-coordinates of a similar flock in figure 3. The two states of
flight, horizontal motion of the flock at a height around h and the landing are
clearly visible. To see whether the landings are synchronised or not we have
plotted the percentage of the landed birds in the flock as a function of time,
in the presence of coupling, in the absence of coupling (f a
z = 0) and in a mean
field case (meaning the radius of N is infinity, not R) in figure 4.
As we can see, the transition in the coupled case is much sharper than in the
8
(a) β = 0
(b) β = 0.03
(c) β = 0.05
(d) β = 0.07
Fig. 2. The tracks of the birds projected onto the xy-plane with different levels of
horizontal noise. The different colours represent different birds, where 10% of the
birds were plotted with longer tracks and the rest of the birds with just the end of
their paths. From (a) to (d) we observe a transitions from an unrealistic straight line
of movement, through the biologically relevant phase to the total loss of coherent
movement.
Fig. 3. The z-coordinates of the different birds against time. The flight around
the height h gives way to collective landing over time as more and more birds are
increasingly biased towards moving downward. The parameters were chosen as of
table 1.
uncoupled case, showing considerable synchronisation among the birds. It is
also notable, that increasing the radius of interaction to infinity does not make
the landing process relevantly sharper, it merely decreases the time needed to
make the decision to land. Coherent landing arises from the interplay of three
9
Fig. 4. The percentage of landed birds as a function of time. The red curve corre-
sponds to the case when coupling between the birds is absent, i.e., f a
z = 0, the green
one corresponds to the coupled case, while the blue curve is the mean field case, i.e.,
where the radius of N is infinity instead of R. It is clearly seen that in the presence
of coupling, the landing is much sharper viz. the synchronisation among the birds is
much greater. It is also notable, that increasing the radius of interaction to infinity
does not make the landing process relevantly sharper, it merely decreases the time
needed to make the decision to land. The parameters were chosen as of table 1.
Table 2
The normalised quantities describing the coherence in space and time, with the
parameter values from table 1. T 0
60 is the value of T60 for the zero coupling case and
σ0
xy is the values of σxy during flight.
Quantity
Value
σL/σ
T60/T 0
60
σ0
xy/σxy
0.387 ± 0.007
0.288 ± 0.007
0.639 ± 0.017
forces: f h
z , the vertical noise and the averaging force. The magnitude of the
bias in the noise increases with time to ultimately overcome f h
z , but due to
the averaging force the individual biases are, in a sense, averaged over the
neighbouring birds. Thus for an individual bird it becomes "harder" to land
when the bulk of the flock would still want to fly, and becomes "easier" when
the latter wants to land.
In table 2 we have summarised the quantitative analysis of the landing. For
this we measured the quantities introduced in section 2.3 in around hundred
independent runs of the simulation, and averaged over the obtained data. Note
that the normalised quantities describing temporal coherence show consider-
able synchronisation in the flock. We stress that this happened while a single
bird had information about only a few neighbours, and not the whole flock.
10
Fig. 5. The values of the most important quantities describing the collective landing
for different number of birds, while keeping the density of the birds constant with
the changing of diameter D (all other parameters are as of table 1). Note that
T60/T 0
60 is well below unity and this property persists for much larger flocks than
the one used in most of our measurements.
In figures 5, 6 and 7 we plot the quantities mentioned in table 2, as functions
of the number of birds in the flock, the standard deviation of the distribution
of ti-s, and the magnitude of the vertical noise, respectively. While changing
the number of birds in the flock, we keep the density of the birds constant by
appropriately changing the horizontal diameter D of the flock. Note that our
previous statements about temporal coherence hold for considerable changes
of these attributes of the flock.
In figure 5 we see that decreasing the number of birds, while keeping the ratio
of the number of birds and the volume set by the boundaries of the forces f h
z
and f c
xy a constant, makes the flock less coherent. This is due to the fact that
with the decrease in the number of birds, the fluctuations in the number of
neighbours, with respect to the individual birds, become important. In figure 6
we see that for small σ-s the value of σL/σ grows well above unity. This
is because there is an inherent difference in landing times, due to the birds
starting from different heights and thus arriving at the ground at different
times, and also due to the fact that repulsion prevents the birds from landing
on top of each other. The effect of this inherent difference becomes noticeable
compared to low σ values.
Figure 7 shows that for any given coefficient of f h
z and C, there is maximum
of temporal coherence or, alternatively a minima of T60/T 0
60 and σL/σ, as a
function of the noise. The interplay between f h
z and the vertical noise be-
comes apparent from figure 7b. The location of the minima as a function of
α (the coefficient of the vertical noise) shifts to right, and its value increases
11
Fig. 6. The values of the most important quantities describing the collective landing
for different standard deviation of the depletion times (all other parameters are as
of table 1). Note that T60/T 0
60 is well below unity. For small values of σ, the ratio
σL/σ grows considerably due to the fact, that factors such as the size of the birds
and the height of the flock, become important in deciding the time taken by the
whole flock to land.
(in σL/σ) with the increase of the coefficient of f h
z . To explain this, let us
increase α from zero as we keep C a constant. At small values of α we see
that there is no landing, as the noise is not strong enough to overcome f h
z .
When α is larger (right side of figure 7b) the temporal coherence of the flock
decreases with the increase of α, as one would intuitively think. Between these
two there is regime with a non-trivial minimum of σL/σ (maximum of tem-
poral coherence), where the increase in the magnitude of the noise actually
increases temporal coherence within the flock (see e.g. noise-induced ordering
[24], stochastic resonance [25]). The increase of C decreases the maximum of
temporal coherence the flock can reach and increases the magnitude of noise
needed to reach this maximum.
Our model, naturally, uses assumptions which are results of simplifications
of the true complexity of a flock of birds. We assume local interactions (that
these dominate the landing process) in addition to the a priori assigned intent
of the birds. A possibility would be to include the reaction of a single bird
to the behaviour of the whole flock (taking this into account would need a
number of further arbitrary assumptions), but we have restricted the model
to mostly local interactions, with one bird interacting with an estimated 8-12
birds on average.
12
(a)
(b)
Fig. 7. (a) The values of the most important quantities describing the collective
landing for different magnitudes of the vertical noise (all other parameters are as of
table 1). For values of α slightly less than the smallest shown on the graph no landing
occurs. (b) The temporal coherence of the flock as a function of the magnitude of
the vertical noise and f h
z (all other parameters are as of table 1). The place of the
minimum in α shifts to the right, and the value of the minimum increases as f h
z
becomes stronger. For values of α slightly less than the smallest shown on the graph
no landing occurs.
3 Conclusion
In this report, we have investigated a model where the concepts of collective
decision making and collective motion are intertwined. In particular, we have
introduced a simple phenomenological model for the landing of a flock of birds.
In the model birds are only influenced by the dynamical variables, like position
and velocity, of other birds in their immediate neighbourhood. Heterogeneity
is introduced in the flock in terms of a depletion time after which a bird feels
an increasing bias to move towards the landing surface. The stochastic nature
of the bias ensures that the external and internal effects influencing the flock
are also included. Through our model we have demonstrated a mechanism by
which animals in large group can arrive at an egalitarian decision about the
time of switching from one activity to another in the absence of a leader. Our
results suggest that the coherence of the collective action of landing can be
enhanced by the random perturbations.
Acknowledgement
This research was supported by the EU ERC COLLMOT project.
13
References
[1] L. Conradt and T. J. Roper, Nature 421 (2003) 155.
[2] C. P. Chamley, Rational Herds, Cambridge University Press (2004), Cambridge.
[3] S. Garnier, J. Gautrais, M. Asadpour, C. Jost and G. Theraulaz, Adaptive
Behaviour 17 (2009) 109.
[4] A.J.W. Ward, D. J. T. Sumpter, I. D. Couzin, P. J. B. Hart, J. Krause, PNAS
105 (2008) 6948.
[5] C. Detrain and J.-L. Deneubourg, Advances in Insect Physiology 35 (2008) 123.
[6] Q. Michard and J.-P. Bouchaud, Eur. Phys. J. B 47 (2005) 151.
[7] C. Castellano, S. Fortunato and V. Loreto, Rev. Mod. Phys. 81 (2009) 591.
[8]
I. D. Couzin, J. Krause, R. James, G. D. Ruxton and N. R. Franks, J. theor.
biol. (2002) 218 1.
[9] G. Gr´egoire and H.Chat´e, Phys. Rev. Lett. 92 (2004) 025702.
[10] T. Vicsek, A. Czir´ok, E. Ben-Jacob, I. Cohen and O. Shochet, Phys. Rev. Lett.
75 (1995) 1226.
[11] L. Conradt and T. J. Roper, Proc Royal Soc.B. 274 (2007) 2317.
[12] P. K. Visscher and T. D. Seeley, Behav. Ecol. Sociobiol. 61 (2007) 1615.
[13] J. Gautrais, P. Michelena, A. Sibbald, R. Bon, J.-L. Deneubourg, Anim. Behav.
74 (2007) 1443.
[14] H. Meunier, J.-L. Deneubourg and O. Petit, Primates 49 (2008) 26.
[15] A. Ramseyer, O. Petit and B. Thierry, Behaviour 146 (2009) 351.
[16] J. D. Halley and M. Burd,Insect Soc. 51 (2004) 226.
[17] I. Daruka, Proc. R. Soc. B 276 (2009) 911.
[18] D. Helbing and P. Moln´ar, Phys. Rev. E. 51 (1995) 4282.
[19] K. Bhattacharya and T. Vicsek, New J. Phys. 12 (2010) 093019.
[20] I. Aoki, B. Jpn. Soc. Sci. Fish 48 (1982) 1081.
[21] C. W. Reynolds, SIGGRAPH '87 Conference Proceedings 21 (1987) 25.
[22] E. Sirot, Anim. Behav. 72 (2006) 373.
[23] M. Ballerini, N. Cabibbo, R. Candelier, A. Cavagna, E. Cisbani, I. Giardina, V.
Lecomte, A. Orlandi, G. Parisi, A. Procaccini, M. Viale, V. Zdravkovic, Proc.
Natl. Acad. Sci. USA 150 (2008) 1232.
14
[24] L. Gammaitoni, P. Hanggi, P. Jung, F. Marchesoni, Rev. Mod Phys. 70 (1998)
223.
[25] C. Van der Broeck, J. M. R. Parrondo, R. Toral, R. Kawai, Phys. Rev. E 55
(1997) 4084.
15
|
1609.05202 | 3 | 1609 | 2017-05-28T06:20:21 | A versatile framework for simulating the dynamic mechanical structure of cytoskeletal networks | [
"physics.bio-ph"
] | Computer simulations can aid in understanding how collective materials properties emerge from interactions between simple constituents. Here, we introduce a coarse-grained model that enables simulation of networks of actin filaments, myosin motors, and crosslinking proteins at biologically relevant time and length scales. We demonstrate that the model qualitatively and quantitatively captures a suite of trends observed experimentally, including the statistics of filament fluctuations, mechanical responses to shear, motor motilities, and network rearrangements. We use the simulation to predict the viscoelastic scaling behavior of crosslinked actin networks, characterize the trajectories of actin in a myosin motility assay, and develop order parameters to measure contractility of a simulated actin network. The model can thus serve as a platform for interpretation and design of cytoskeletal materials experiments, as well as for further development of simulations incorporating active elements. | physics.bio-ph | physics | A versatile framework for simulating the dynamic
mechanical structure of cytoskeletal networks
S. L. Freedman, S. Banerjee, G. M. Hocky, A. R. Dinner
Abstract
Computer simulations can aid in understanding how collective materials properties emerge
from interactions between simple constituents. Here, we introduce a coarse-grained model that
enables simulation of networks of actin filaments, myosin motors, and crosslinking proteins
at biologically relevant time and length scales. We demonstrate that the model qualitatively
and quantitatively captures a suite of trends observed experimentally, including the statistics of
filament fluctuations, mechanical responses to shear, motor motilities, and network rearrange-
ments. We use the simulation to predict the viscoelastic scaling behavior of crosslinked actin
networks, characterize the trajectories of actin in a myosin motility assay, and develop order
parameters to measure contractility of a simulated actin network. The model can thus serve as
a platform for interpretation and design of cytoskeletal materials experiments, as well as for
further development of simulations incorporating active elements.
7
1
0
2
y
a
M
8
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
3
v
2
0
2
5
0
.
9
0
6
1
:
v
i
X
r
a
1
1
Introduction
The actin cytoskeleton is a network of proteins that enables cells to control their shapes, exert forces
internally and externally, and direct their movements. Globular actin proteins (G-actin) polymerize
into polar filaments (F-actin) that are microns long and nanometers thick. Many different proteins
bind to actin filaments; such proteins often have multiple binding sites that enable them to crosslink
actin filaments into networks that can transmit force. Myosin proteins are composed of head, neck,
and tail domains and aggregate via their tails to form minifilaments that can attach multiple heads
to actin filaments (1). Each myosin head can bind to actin and harness the energy from ATP
hydrolysis such that a minifilament can walk along an actin filament in a directed fashion-i.e., it
is a motor. These dynamics have been extensively studied, and it is well understood, for example,
how they give rise to muscle contraction. In muscle cells, myosin II minifilaments bind to regularly
arrayed antiparallel actin filaments and walk toward the barbed ends (2). In other types of cells
lacking this level of network organization, however, the ways in which the elementary molecular
dynamics act in concert to give rise to complex cytoskeletal behaviors remain poorly understood.
Addressing this issue requires a combination of experiment, physical theory, and accurate sim-
ulation. The last of these is our focus here-we present a nonequilibrium molecular dynamics
framework that can be used to efficiently explore the structural and dynamical state space of as-
semblies of semiflexible filaments, molecular motors, and crosslinkers. By allowing independent
manipulation of parameters normally coupled in experiment, this computational model can guide
our understanding of the relationship between the microscopic biochemical protein-protein inter-
actions and the macroscopic mechanical functions of assemblies. Additionally, because the model
simulates filaments, motors, and crosslinkers explicitly, we can elucidate microscopic mechanisms
by studying its stochastic trajectories at levels of detail that are experimentally inaccessible. The
fact that complex behaviors can emerge from simple interactions also allows simulations to be used
to evaluate predictions from theory.
In this work we detail the model and demonstrate that, it reproduces an array of known ex-
perimental results for actin filaments, assemblies of actin and crosslinkers (passive networks), and
assemblies of actin and myosin (active networks). We go further to provide new experimentally
testable predictions about these systems. For single polymers, we reproduce the spatiotemporal
fluctuation statistics of actin filaments. For passively crosslinked networks, we reproduce known
stress-strain relationships and predict the dependence of the shear modulus on crosslinker stiffness.
For active networks, we reproduce velocity distributions of actin filaments in myosin motility as-
says and show how one can tune their dynamical properties by varying experimentally control-
lable parameters.
In separate studies, we use the model to clarify microscopic mechanisms of
actomyosin contractility and investigate how assemblies of actin filaments and crosslinkers can be
tunably rearranged by myosin motors to form structures with distinct biophysical and mechani-
cal functions (3). The collection of this benchmark suite is itself useful, as prior models (4–14)
have focused on specific cytoskeletal features, making the tradeoffs needed to capture the selected
behaviors unclear.
Indeed, our model builds on earlier studies, which we briefly review to make clear similarities
and differences of the models (see also (15) for a list of cytoskeletal simulations). The most finely
detailed simulations focus on the motion of a myosin minifilament with respect to a single actin
filament. Erdmann and Schwarz (9) used Monte Carlo simulations to verify a master equation
describing the attachment of a minifilament and, in turn, the duty ratio and force velocity curves
2
as functions of the myosin assembly size. Stam et al. (16) used simulations to study force buildup
on a single filament by a multi-headed motor and found distinct timescale regimes over which
different biological motors could exert force and act as crosslinkers. These models of actin-myosin
interactions are important for understanding the mechanics at the level of a single filament, and
their results can be incorporated into larger network simulations.
A number of publications have dealt with understanding the rheological properties of cross-
linked actin networks (4–7). For example, to study the viscoelasticity of passive networks, Head
et al. (5) distributed filaments randomly on a two-dimensional (2D) plane, crosslinked filament
intersections to form a force-propagating network, sheared this network, and let it relax to an
energy minimum. From this model, they were able to identify three elastic regimes that were
characterized by the mean distance between crosslinkers and the temperature. Dasanyake and
coworkers (8) extended this model to include a term in the potential energy that corresponded to
myosin motor activity and observed the emergence of force chains that transmit stress throughout
the network. These studies address questions about how forces propagate and how crosslinker
densities alter mesh stiffness.
Other studies characterized network structure and contractility as functions of model parame-
ters. Wang and Wolynes (10) considered a graph of crosslinkers (nodes) and rigid filaments (edges)
in which motor activity was simulated via antisymmetric kicks along the filaments. They calculated
a phase diagram for contractility as a function of crosslinker and motor densities. Such a simu-
lation can provide qualitative insights into general principles of filament networks, but the model
did not account for filament bending, and structures were sampled via a Monte Carlo scheme that
was not calibrated to yield information about dynamics. Cyron et. al. (17) used Brownian dynam-
ics simulations to investigate structures that can form via mixtures of semiflexible filaments and
crosslinkers and determined a phase diagram and phase transitions (18) between differently bun-
dled actin networks that form as one varies crosslinker density and crosslinker-filament binding
angle. Nedelec and coworkers performed dynamic simulations of assemblies of semiflexible mi-
crotubules and kinesin motor proteins, which share features with assemblies of F-actin and myosin;
they used their simulation package, CytoSim, to understand aster and network formation in micro-
tubule assays (11) and showed recently that the model can be adapted to treat actin networks (12).
Gordon et al. (13), Kim (14), and most recently Popov et al. (15) similarly simulated dynamics of
F-actin networks and included semiflexible filaments, motors and crosslinkers. By varying motor
and crosslinker concentrations, Gordon et al. (13) and Popov et al. (15) showed various structures
that can emerge from assemblies of this type, and Kim (7, 14) additionally quantified how these
changes could effect force propagation within the network.
We have strived to include many of the best features of these preceding models in our model.
We use the potential energy of Head et al. (5) for filament bending and stretching. However, in
contrast to (5, 8), which simply relax the network, we simulate the stochastic dynamics, including
thermal fluctuations, crosslinkers and motors binding and unbinding, and the processive activity of
myosin. The force propagation rules and kinetic equations for binding and unbinding are similar
to those of (11, 13), while the length and time scales simulated are on the order of those performed
in (12, 14). We expand on these works by combining and documenting key elements in a single
model, demonstrating that the model can capture experimentally determined trends for cytoskeletal
materials quantitatively, and illustrating how the model can be used to study systems of current
experimental interest.
3
2 Materials and Methods
To access the time and length scales relevant to cytoskeletal network reorganization, we treat actin
filaments, myosin minifilaments, and crosslinkers as coarse-grained entities (Fig. 1A). We model
actin filaments as polar worm-like chains (WLC) such that one end of the WLC represents the
barbed end of an actin filament and the other represents the pointed end. We model crosslinkers
as Hookean springs with ends that can bind and unbind from filaments. Thus, the connectivity
of a network and, in turn, its capacity for force propagation varies during simulations. We model
molecular motors similarly to crosslinkers except that each bound motor head can walk toward
the filament barbed end with a load-dependent speed. The motors can slide filaments, translocate
across filaments, and increase network connectivity. We simulate the system using Langevin dy-
namics in 2D because the in vitro experiments we wish to interpret are quasi-two-dimensional, and
approximating the system as 2D allows us to treat larger systems for longer times. To account for
the fact that a three-dimensional (3D) system would have greater conformational freedom, we do
not include steric interactions for our filaments, motors and crosslinkers. This implementation of
filaments, motors and crosslinkers, which we detail below, allows for motor-driven filament sliding
and filament buckling, as seen in Fig. 1D-E. A complete list of model parameters, their values, and
references is provided in Table 1.
Figure 1: Overview of the model. See Materials and Methods for details. (A) Schematic of a configuration
of the model. Filaments are red, crosslinkers are green, and motors are black. (B) Expanded view of the
actin filament representation: a chain of beads connected by springs with spring constant ka, rest length la,
and bending modulus κB, as detailed in Filaments. (C) The process by which a crosslinker finds a filament
to bind, as detailed in Crosslinkers. The solid red link is indexed to the grid points marked with either red
or purple stars, and the solid green motor head searches the grid points marked with either green or purple
stars for links to bind. The crosslinker head then stochastically binds to the nearest spot on the filament (see
Section S1 and Fig. S1A in the Supporting Materials) here marked with a purple ×. (D) Successive images
of two antiparallel 10 µm filaments (barbed end marked in blue) interacting with one motor at the center.
The motor binds to both filaments and slides them past each other. (E) Similar to (D) but with a crosslinker
that pins the top filament's pointed end to the bottom filament's barbed end. The motor, bound to both,
walks toward the barbed end of the bottom filament and buckles the top filament.
2.1 Filaments
The WLC model for actin filaments is implemented as a chain of N + 1 beads connected by N
harmonic springs (links) and N − 1 angular harmonic springs, as depicted in Fig. 1B. The N
4
ka, la, κBθi-1θiθi+1rc3s0s10sDEACBkxlonkxloffv0linear springs penalize stretching and keep the filament's average end-to-end length approximately
constant. The N − 1 angular springs penalize bending and determine the persistence length for a
free filament. The filament configurations are governed by the potential energy Uf:
Uf = U stretch
f
f
+ U bend
((cid:126)ri − (cid:126)ri−1 − la)2
U stretch
f
=
U bend
f =
ka
2
κB
2la
N(cid:88)
N(cid:88)
i=1
i=2
θ2
i ,
(1)
where (cid:126)ri is the position of the ith bead on a filament, θi is the angle between the ith and (i − 1)th
links, ka is the stretching force constant, κB is the bending modulus, and la is the equilibrium
length of a link. In practice, Uf enters the simulation through its Cartesian spatial derivatives (i.e.,
the forces in Eq. 6). In this regard, it is important to note that linearized forms for the bend-
ing forces are employed in the literature for filaments whose length is constrained via Lagrange
multipliers (11), but we found that it was necessary to use the full nonlinear force to obtain consis-
tent estimates for the persistence length, Lp, for bead-spring-chain filaments (see Actin filaments
exhibit predicted spatial and temporal fluctuations, below). We thus employ the full nonlinear
Cartesian forces throughout this work, using the expressions in Appendix C of (19) following the
implementation in the LAMMPS Molecular Dynamics Simulator (20).
The bending force constant is derived from the persistence length Lp such that κB = LpkBT ,
where kB is Boltzmann's constant, and T is the temperature (21). Experimentally, Lp = 17 µm, so
κB = 0.068 pNµm2 for T = 300 K (22).
The elasticity per unit length measured for actin filaments with lengths on the order of a micron
is 55 ± 15 pN/nm (23, 24). This implies that a reasonable value for the segment stretching force
constant, ka, would be of this order of magnitude. However, simulating a network of such stiff
filaments is computationally infeasible since the maximum timestep of a simulation is inversely
proportional to the largest force constant in the simulation (25). Therefore, we set ka to a smaller
value than estimated from experiment. We note that prevalent extensile behavior, which occurs
when filaments interact with two populations of motors with opposite polarities (26), would neces-
sitate using a more realistic ka. However, because ka (cid:29) κB/l3
a still, upon compression, filaments
prefer bending to stretching, and, as we show, the ability of our model to capture contractile net-
work properties quantitatively is not compromised. Unless otherwise indicated, we use la = 1 µm,
because it is the largest segment length that results in the expected spatial and temporal fluctuations
for filaments (see Actin filaments exhibit predicted spatial and temporal fluctuations, below). We
note that very high motor densities can cause filaments to buckle at length scales of ∼ 1 µm, and
in these cases it would be necessary to use a smaller la to capture those effects (27).
2.2 Crosslinkers
There are a variety of different actin binding proteins that serve as crosslinkers in the cell cortex,
including filamin, fascin, and α-actinin. Crosslinkers connect filaments dynamically and propagate
force within the network. Thus, the crosslinkers in our model must be able to attach and detach
from filaments with realistic kinetic rules and be compliant when bound. To this end, we model
5
them as Hookean springs with stiffness kxl and rest length lxl. Like actin filaments, the Young's
modulus of most crosslinkers is significantly higher than would be reasonable to simulate; there-
fore, for network simulations without large external forces, we set kxl = ka so that the bending
mode of actin filaments is significantly softer than the stretching mode of crosslinkers. The rest
length lxl corresponds to the size of the crosslinker and therefore differs based on the particular
actin binding protein one wishes to study.
The statistics of the bound (on) and unbound (off ) states of each crosslinker are determined by
a potential energy of the form
Uxl = U stretch
xl
+ U bind
xl
(I1 + I2)
U stretch
xl
=
1
2
kxl((cid:126)r1 − (cid:126)r2 − lxl)2
xl /kof f
kon
xl
xl = −kBT ln
U bind
(cid:16)
(cid:17)
(2)
where (cid:126)r1(2) is the position of head 1(2), I1(2) is 1 if head 1(2) is bound and 0 otherwise, and kon
xl
(kof f
xl ) are the rates of binding (unbinding).
Owing to the form of Eq. 2 and the Monte Carlo rule for binding (below), it is inefficient for a
crosslinker to attempt attachment to every filament link in the simulation box. Rather, we assign a
cutoff distance rc =(cid:112)kBT /kxl such that if the distance between a motor and a filament is greater
than rc the probability of attachment is zero. This implementation allows us to use the following
neighbor list scheme, illustrated in Fig. 1C, to determine crosslinker-filament attachment. A grid
of lattice size of at least rc is drawn on the 2D plane of the simulation, and each filament link is
indexed to the smallest rectangle of grid points that completely enclose it. In practice the lattice
size is generally larger than rc due to memory constraints, and is denoted by the model parameter
g, the number of grid points per µm in both the x and y directions. Since a crosslinker head cannot
bind to a filament link that is farther away than rc, it suffices for a crosslinker head to only attempt
attachment to the nearby filament links indexed to its four nearest grid points.
At each timestep of duration ∆t, we enumerate the accessible filament links available to each
unbound head. For each link, we determine the nearest point to the head's present position and
compute a Metropolis factor for moving to that point: P of f→on
/kBT )]
and stays
(28). The head then binds to accessible filament link i with probability (kon
= min[1, exp(−∆U stretch
xl ∆t)P of f→on
unbound with probability 1 −(cid:80)
(29).
xl ∆t)P of f→on
xl,i
i(kon
At each timestep, we attempt to move each bound head to a position (cid:126)ru generated by reversing
the displacement made upon binding, rotated to account for filament reorientation in the inter-
vening time. This choice of (cid:126)ru allows us to satisfy detailed balance for binding and unbinding
by accepting the unbinding transition with probability (kof f
/kBT )], as
explained in Section S1 in the Supporting Materials.
xl ∆t) min[1, exp(−∆U stretch
xl
xl,i
xl,i
xl,i
When both crosslinker heads are attached to filaments, the crosslinker is generally stretched
or compressed. We propagate the tensile force stored in the crosslinker onto the filaments via the
lever rule described in (13, 30). Specifically, if the tensile force of a crosslinker head at position
(cid:126)rxl between filament beads i and i + 1 is (cid:126)Fxl, then,
(cid:126)ri+1 − (cid:126)rxl
(cid:126)ri+1 − (cid:126)ri
(cid:126)Fi = (cid:126)Fxl
(cid:126)Fi+1 = (cid:126)Fxl − (cid:126)Fi
6
(3)
are the forces on beads i and i + 1 respectively due to the crosslinker.
2.3 Motors
In the present work, we focus on the motor protein myosin II. As mentioned above, tens of myosin
II proteins aggregate into bipolar assemblies called myosin minifilaments (16). For both myosin
minifilaments, and monomeric myosin, motility assay experiments have shown that, on aver-
age, bound myosin heads walk toward the barbed end of actin filaments at speeds in the range
0.2 − 4 µm/s (31–34). Since myosin also functions to increase the local elasticity of networks
where it is bound, we model a motor similarly to a crosslinker, in that it behaves like a Hookean
spring with two heads, a stiffness km, and a rest length lm. The two heads of this spring do not cor-
respond directly to individual myosin protein heads; rather each of them represents tens of myosin
molecules. Experimentally minifilaments have a very high Young's modulus, and it is unlikely that
their lengths change noticeably in cytoskeletal networks. As with the passive crosslinkers, we set
km = ka so that filament bending is still the softest mode. The rest length was set to the average
length of minifilaments, lm = 0.5 µm (1). Attachment and detachment kinetics, as well as force
propagation rules for motors, are the same as for crosslinkers, subscripted with m instead of xl in
Eqs. (2) and (3).
Unlike crosslinkers, motors move towards the barbed end of actin filaments to which they
are bound at speeds that decrease with tensile force along the motor. Myosin motors have been
observed to stop walking when the force on them exceeds the stall force, Fs ≈ 4 pN, and most do
not step backward (35, 36). We model this behavior by giving each motor head a positive velocity
in the direction of the barbed end of the filament to which it is attached; this velocity linearly
decreases with the motor's tension projected on the filament, i.e.,
(cid:40)
(cid:41)
(cid:126)Fm · r
Fs
v( (cid:126)Fm) = v0 max
1 +
, 0
,
(4)
where v0 is the unloaded motor speed, Fm = −k((cid:126)r1 − (cid:126)r2 − lm) is the spring force on the motor,
and r is the tangent to the filament at the point where the motor is bound; r points toward the
pointed end of the filament. In the simulations below, we use a value of v0 = 1 µm/s, which is
within range of experimental measurements, but we use a lower value of Fs = 0.5 pN, so that
motors are not stretched to unphysical lengths as they walk.
m . We found that kend
If the length of a motor's step is larger than the remaining length of filament, then the myosin
moves to the barbed end of the filament. At the barbed end, it has speed v0 = 0, and detachment
rate kend
m yielded reasonable results for motility assay and contractile
network simulations. In experiments, where each myosin minifilament contains many myosins, a
lower barbed end affinity may arise from fewer of the minifilament's myosins remaining attached
to the actin filament. In the program, we treat crosslinkers and motors with equivalent objects, but
set v0 = 0, and kend
for the crosslinkers.
m = 10kof f
xl = kof f
xl
2.4 Dynamics
We use overdamped Langevin dynamics to solve for the motion of filament beads, motors, and
crosslinkers. The Langevin equation of motion for a spherical bead of mass m, radius R at position
7
(cid:126)r(t) at time t, forced by (cid:126)F ((cid:126)r(t)) in a medium with dynamic viscosity ν is
m(cid:126)r(t) = (cid:126)F ((cid:126)r(t)) + (cid:126)B(t) − (cid:126)r(t)/µ,
(5)
where (cid:126)B(t) is a Brownian forcing term that introduces thermal energy, and we use the Stokes
relation µ = (6πRν)−1 in the damping term. The fastest motions in this simulation are the
filament bead fluctuations. Taking the bead radius to be 0.5 µm, the maximum speed to be
(2kBT µ/∆t)1/2 = 200 µm/s and the dynamic viscosity to be ν = 0.001 Pa·s (corresponding
to water), the Reynolds number is very low: Re ≈ 10−4. Hence, we treat the dynamics as
overdamped and set m = 0 in Eq. 5. Furthermore, in the limit of small ∆t, we may write
(cid:126)r(t) ≈ ((cid:126)r(t + ∆t) − (cid:126)r(t))/∆t. These two approximations allow us to rewrite Eq. 5 as
(cid:126)r(t + ∆t) = (cid:126)r(t) + (cid:126)F ((cid:126)r(t))µ∆t + (cid:126)B(t)µ∆t.
For the Brownian term, we use the form of Leimkuhler and Matthews (37):
(cid:115)
(cid:32) (cid:126)W (t) + (cid:126)W (t − ∆t)
(cid:33)
,
(cid:126)B(t) =
2kBT
µ∆t
2
(6)
(7)
where (cid:126)W (t) is a vector of IID random numbers drawn from the standard normal distribution. This
numerical integrator minimizes deviations from canonical averages in harmonic systems; given
that all the mechanical forces in our model are harmonic, we expect this choice to yield accurate
statistics in the present context as well. The value for ∆t in Eq. 6 is most strongly dependent on
the largest force constant in the simulation, ka, but also depends on other simulation parameters
for both motors and crosslinkers, such as v0, kon, and kof f. Table 1 can be used as a rough guide
for how high one can set the value of ∆t for a given set of input parameters; e.g., for a contracting
m = 0.1 s−1, and
network with ka = 1 pN/µm, v0 = 1 µm/s, kon
m = 10 s−1, a value of ∆t = 2 × 10−5 s is just low enough to iteratively solve Eq. 6 without
kend
accumulating large errors.
m = 1 s−1, kof f
xl = kof f
xl = kon
m = kend
2.5 Environment
In general we use periodic boundary conditions so as to limit finite-size effects. We implemented
square boundaries to model closed systems, as well as Lees-Edwards boundaries for shearing sim-
ulations (19). The dimensions of the simulation box (Table 1) were chosen to be five times the
contour length of filaments so as to be large enough to avoid artifacts due to the self-interaction of
constituent components.
To ignore steric interactions, the fraction φ = Nf π(D/2)2L/V of Nf actin filaments (length
L and diameter D) in a volume V must be lower than the critical volume fraction at which steric
interactions yield an isotropic to nematic transition, which for long worm-like chains (D (cid:28) Lp
and D (cid:28) L) is φc = 5.4D/L (38, 39). For a network of 500 filaments of length L = 10 µm
and diameter D = 0.01 µm, in a 50 × 50 × 0.1 µm3 plate, this condition is fulfilled, since φ =
0.0015 < φc = 0.0054. While it is difficult to estimate the exact thickness of in vitro experimental
actomyosin assays due to the complexity of their preparation, we estimate that they are not thinner
than 0.1 µm (40). We have also ignored hydrodynamic interactions between filament beads; the
restriction to low packing fraction obviates the need to incorporate anisotropic drag, so we take µ
to be equivalent for both transverse and longitudinal motion (41).
8
3
Implementation
is
implemented as
that
an open source C++ package
called Active Fila-
The model
is available for download at http://dinner-
ment Network Simulation (AFiNeS)
group.uchicago.edu/downloads.html.
Installation instructions are available in the README
file in the top directory of the AFiNeS package, and all information needed to reproduce the
materials in this paper are available in the subfolder "versatile framework paper". To run a
simulation, a user must compile the code into an executable (e.g., with the provided Makefile)
and create an output directory. A user can set parameters using command line arguments or
a file. For example, if the user has compiled the code into the executable "afines", created
the output directory "test", and wants to run a simulation of 500 10 µm actin filaments (with
la = 1 µm), interacting with 0.2 motors/µm2, and 1 crosslinker/µm2 (passive motors), in a cell
that is 50 µm × 50 µm, for 100 s, he or she could write the following to the file my config.cfg
xrange=50 # system size in X
yrange=50 # system size in Y
npolymer=500 # number of actin filaments
nmonomer=11 # number of actin beads per filament
a_motor_density=1 # motor density
p_motor_density=1 # crosslinker density
tf=100 # duration of simulation
dir=''test'' # output directory
and then run the code using the command
afines -c my_config.cfg
Alternatively, the user could bypass the configuration file and issue the following command:
afines --xrange 50 --yrange 50 --npolymer 500 --nmonomer 11 \
--a_motor_density 1 --p_motor_density 1 --tf 100 --dir test
In this example, all other parameters were set to their default values (see README file for full list
of program parameters). With an executable compiled using g++ with the -O3 optimization flag
and run on an Intel E5-2680 node with 2 Gb of memory and a 2.7 GHz processor, this example
required less than 1.5 days of wall-clock time. In general, the wall-clock time of the simulation
scales linearly with system size (Fig. 2).
9
Figure 2: Wall clock time for a 10000-step simulation with step size ∆t = 0.0001 s. (A) For a constant
system size, run time scales linearly or sublinearly as both filament density (red dots) and motor density
(blue dots) are increased independently. If both are increased together (black dots), a quadratic scaling is
approached for large numbers of particles. (B) Blue: At constant motor, filament, and grid densities, run
time scales linearly with system size (i.e., the area of the simulation box, XY ). Red: At constant system size,
run time decreases with increasing grid density, g2, and thereby the number of neighbor-list grid elements,
g2XY , used to calculate motor-filament interactions. All benchmarks are for an Intel E5-2680 node with 2
Gb of memory and a 2.70 GHz processor.
10
ttt-1t2ABSymbol Description (units) (references)
Simulation
Table 1: Parameter Values
Actin Filaments
Number of filaments
Number of beads per filament
Link rest length (µm)
Stretching force constant (pN/µm)
Bending modulus (pNµm2) (22)
Myosin Motors
Motor density (µm−2)
Rest length (µm) (1)
Stiffness (pN/µm)
Max attachment rate (s−1)
Max detachment rate (s−1)
Max detachment rate at barbed end
(s−1)
Unloaded speed (µm/s) (31)
Stall force of myosin (pN) (42)
Crosslinkers
Crosslink density (µm−2)
Rest length (Filamin) (µm) (43)
Stiffness (pN/µm)
Max attachment rate (s−1)
Max detachment rate (s−1)
Environment
Dynamics timestep (s)
Total simulated time (s)
Length and width of assay (µm)
Grid density (µm−1)
Temperature (K)
Dynamic viscosity (Pa·s)
Strain (%) (44)
Time between sequential strains (s)
Nf
NB
la
ka
κB
ρm
lm
km
kon
m
kof f
m
kend
m
v0
Fs
ρxl
lxl
kxl
kon
xl
kof f
xl
∆t
TF
X, Y
g
T
ν
∆γ
trelax
Lp
20
[21, 201]
[0.1, 1]
[0.01, 1000]
[0.005, 800]
Shear
500
11
1
1000
Motility
Assay
Network
1
[2, 26]
1
1
500
11
1
1
0.068
0.068
0.068
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
[10−6, 10−3]
2000
n/a
n/a
300
0.001
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
n/a
0.42
0.150
[0.1, 1000]
1
0.1
10−7
0.5
75
2
300
0.001
0.001
0.001
[0, 9]
0.5
1
[0.001, 2]
1
10
1
0.5
n/a
n/a
n/a
n/a
n/a
0.2
0.5
1
1
0.1
1
1
0.5
1
0.150
1
1
0.1
0.00005
0.00002
1000
50
2
300
0.001
n/a
n/a
400
50
2
300
0.001
n/a
n/a
11
4 Results and Discussion
In this section, we numerically integrate the model to obtain stochastic trajectories and compare
their statistics to known analytical results for semiflexible polymers and networks, as well as ex-
perimental observations. We also use the model to investigate these systems, including how the
viscoelasticity of semiflexible polymer networks depends on crosslinker stiffness, and how the
extent of directed motion in actin motility assays depends on filament and motor characteristics.
Finally, we use the model to show how one can quantify contractility in a simulated actin network.
4.1 Actin filaments exhibit predicted spatial and temporal fluctuations
The persistence length of a semiflexible filament with bending modulus κB is expected to be
Lp = κB/kBT . However, when simulating the dynamics, approximations can enter both the eval-
uation of the forces and the discretized numerical integration of the equations of motion. Because
the persistence length is a measure of filament bending fluctuations, and not an input to the simu-
lation, its dependence on simulation parameters must be determined numerically. As discussed in
Filaments and further below, some care is required to obtain reliable estimates of Lp.
links i and i − 1 of an N link chain results in a local change in free energy of (κB/2la)θ2
For a two dimensional filament it is possible to show analytically that if a small bend between
i , then
(cid:104)θ2(l)(cid:105) = l/Lp
(8)
(cid:104)cos(θ(l))(cid:105) = exp (−l/2Lp),
(9)
where θ(l) = θj − θi, l = la(j − i) (2 ≤ i < j ≤ N ) (45). To test our WLC model against
these predictions, we let 20 filaments of L = 20 µm and κB = 0.068 pNµm2 fluctuate at T = 300
K for Tf = 2000 s and measured the resulting filament configurations. The configurations saved
were chosen to be 2 s apart, since the decorrelation time for θ(l) was at most 1.1 s (see Section S2
and Fig. S2 for details). The first 100 s of each simulation was disregarded as filaments had
not yet equilibrated. For each of the 20 filaments, we evaluated (cid:104)θ2(l)(cid:105) and (cid:104)cos(θ(l))(cid:105) for each
l ∈ 1, 2, . . . , 19 µm from its 1900 saved configurations. We show the average for each of these
values over all filaments in Fig. 3B, along with the expected behavior, given the input κB.
12
Figure 3: Spatial and temporal fluctuations of the bead-spring WLC. (A) Schematic of a filament and
the order parameters that characterize its fluctuations. Spatial fluctuations are characterized by the angle
between two tangent vectors (cid:126)ds and (cid:126)ds+l along the filament as a function of the contour length between
them, l. Temporal fluctuations are characterized by the eigenvalues λ1,2(t) of the covariance matrix of
filament endpoint positions as a function of time. The red arrow indicates the larger moment (λ1, measuring
transverse fluctuations) while the blue arrow indicates the smaller moment (λ2, measuring longitudinal
fluctuations). (B) Decorrelation of tangent vectors (red circles) and fluctuations in angles between links
√
(blue circles) as a function of the arc length between them. For the N = 20 filaments analyzed, the blue
(red) dots show the mean of (cid:104)θ(l)2(cid:105) ((cid:104)cos (θ(l))(cid:105)) and the error bars show their standard errors, σ/
N,
where σ is their standard deviation. Dashed lines show expected behavior for κB = 0.068 pNµm2. (C)
Eigenvalues of covariance matrices for the positions of endpoints of filaments as a function of time. Red
dots show λ1(t), which is expected to be proportional to t3/4 (red line) while blue dots show λ2(t), which
is expected to be proportional to t7/8 (blue line). Standard error is smaller than the size of the data points.
As alluded to above, the numerical integration can make the persistence length depend on
simulation parameters in nonobvious ways. Consequently, we measured the sensitivity of Lp to
independent variations of κB, la, and ka. The results shown in Fig. 4 are obtained from using the
definition Lp = 1/(d(cid:104)θ2(l)(cid:105)/dl) (i.e., the inverse of the slope of the "blue" line in 3B). Fig. 4A
shows that in the range of κB ∈ [1, 105] µm×kBT , Lp determined from the simulation agrees well
with the input bending modulus, and can be easily tuned to simulate filaments of varying rigidity.
Fig. 4B shows that for a wide range of link stiffnesses, Lp is independent of ka. We also tested
the dependence of Lp on the link rest length, la. In thermal equilibrium, the variance of the link
lengths is (cid:104)∆l2
a(cid:105) = kBT /ka. Thus, to keep the fluctuations in the filament's contour length L
constant, one should set ka ∝ l−2
a . In practice, this scaling is computationally difficult to achieve
when la < 0.3 µm because high ka requires a very small ∆t in Eq. 6. We therefore used a less
steep variation, ka = 1 pN/la, and show in Fig. 4C that consistent values of Lp are obtained when
la ∈ [0.1, 1] µm. We thus see that, there is a range in which Lp is independent of the filament
link parameters, ka and la, although high stiffness and low link length both require using a small
timestep, and therefore limit the duration of the simulation. In Section S3 (Fig. S4) we measure
the persistence length of fibers simulated using Cytosim and obtain similar results.
13
BCλA√
Figure 4: Dependence of the persistence length on the parameters for numerically integrated semiflexible
filaments. Error bars are σ/
N, where σ is the standard deviation of the values of Lp obtained from fitting
a line to the first 5 data points of (cid:104)θ2(l)(cid:105) for each of the N = 20 filaments simulated. The dashed lines
show the predicted persistence length, based on the input bending modulus κB. The default parameters are
κB = 17µm×kBT , ka = 1 pN/µm, and la = 1µm. In (A)-(C), ∆t ≥ 10−6 s, and the largest ∆t that
yielded stable integration was used. In (C), ka = 1pN/la.
The statistics of temporal fluctuations are also known for semiflexible filaments. Fluctuations
transverse to the filament orientation increase as (cid:104)dr2⊥(cid:105) ∝ t3/4, while longitudinal fluctuations
increase as (cid:104)dr2(cid:105) ∝ t7/8 (46). To determine if our simulations agreed with these theoretical scaling
relations, we followed the procedure outlined in (46) and generated N = 100 initial filament
configurations of a 20 µm filament. This length was chosen because it satisfied the constraint
provided in (46) for the fluctuations of the two ends of the filament to be uncorrelated at long
times (here t = 1 s); i.e., 20 µm > (tkBT /ν)1/8 (κB/kBT )5/8 = 7 µm. For each configuration
we ran M = 1000 simulations of the filament diffusing freely for 1 s. We denote each of the
M positions for each endpoint at each time by (cid:126)re(t). For each of the clouds of points shown
in Fig. 3A, we calculated the moments, as the eigenvalues of the covariance matrix with elements
(cid:104)((cid:126)re(t)·i−(cid:104)(cid:126)re(t)·i(cid:105))((cid:126)re(t)·j−(cid:104)(cid:126)re(t)·j(cid:105))(cid:105) for i, j ∈ {x, y}. The larger eigenvalue λ1(t) corresponds
to the transverse fluctuations (i.e., λ1(t) ∝ t3/4) while the smaller eigenvalue corresponds to the
longitudinal fluctuations (λ2(t) ∝ t7/8). We show these results in Fig. 3C. Each data point is the
average over the 2N M eigenvalues for λ1(t) and λ2(t). As evident, the computed scaling relations
are in good agreement with theoretically predicted behaviors.
4.2 Tunable elastic behavior of crosslinked filament networks
The mechanical properties of crosslinked F-actin have important ramifications for force generation
and propagation within a cell. They are generally inferred from rheological measurements of in
vitro networks (47–50). In a typical experiment, actin and crosslinker proteins are mixed to form a
crosslinked mesh and then sheared in a rheometer by a prestress, σ0. The prestressed network then
undergoes a sinusoidal differential stress of magnitude dσ (cid:28) σ0. By measuring the resulting strain,
14
ABCDone can calculate the differential elastic modulus G(σ0) = dσ/dγ. Results from such experiments
indicate that, in contrast to a purely viscous fluid, crosslinked F-actin networks resist shear, and G
increases nonlinearly with stress indicative of shear stiffening.
In experiments using a stiff crosslinker, such as scruin, the dependence of the differential mod-
ulus on high prestress is G ∝ σ3/2
(47, 50). Force-extension experiments with semiflexible fila-
ments, in which one directly measures the force F required to extend a filament by a distance l,
yield a remarkably similar relationship, dF/dl ∝ F 3/2 (51, 52). As remarked in (47), this sug-
gests that the shear stiffening is a direct result of the nonlinear force-extension relationship of actin.
Rheology studies using more compliant crosslinkers, such as filamin, have found a softer response,
G ∝ σ0, indicating that a significant amount of stress is mediated through the crosslinkers, and not
the filaments (49). These results suggest that the strain stiffening behavior of a crosslinked network
can be tuned by varying the crosslinker stiffness.
0
To test this possibility and benchmark our simulations, we subjected passive networks com-
prised of filaments and crosslinkers to shear. We initialized each simulation with N = 500 ran-
domly oriented filaments of length 15 µm in a square box of area 75 µm × 75 µm. A 0.150 µm
crosslink (corresponding to the length of filamin) was initially placed at each filament intersec-
tion. To inhibit network restructuring, the detachment rate of the crosslinkers was set to zero. We
performed 24 such simulations, each with a different crosslinker stiffness in the range 0.1 − 1000
pN/µm.
Simulating shear rheology experiments requires modifying the equations of motion and the
In general, planar Couette flow can be
boundary conditions to achieve a planar Couette flow.
simulated via molecular dynamics using Eq. (4.1) in (53):
mx = Fint,x + γy
my = Fint,y,
(10)
where x and y are the Cartesian coordinates of a particle being sheared, Fint,x, Fint,y are the internal
forces on those particles and γ is the strain. Simultaneously, the upper and lower boundaries must
be sheared by the total strain on the simulation box (19). Comparing Eq. 10 with Eq. 5, we
substitute Fint,x = F (x(t)) + Bx(t) − x(t)/µ. In the overdamped limit, xi = 0, so implementing
Eq. 10 is equivalent to updating filament bead positions via Eq. 6, and shifting the horizontal
position of a bead (xi) by
(cid:16) yi
(cid:17)
xi → xi + ∆γ
Y
,
(11)
where ∆γ = γ∆t and Y is the simulation cell height. The boundary conditions follow the Lees-
Edwards convention (19).
Since moving the particles ∆γ is equivalent to the addition of a significant external force on the
system, it is necessary to let the network relax for a specified amount of time trelax after each shear
event, before measuring the network's internal energy. The magnitude of trelax depends on ∆γ,
which in turn depends on the chosen discretization of the strain and the timestep ∆t. As shown in
Section S2.2 and Fig. S3, we found that ∆γ = 0.001, ∆t = 10−7 s, and trelax = 0.001 s yielded a
stable planar Couette flow, with high enough strains to observe strain stiffening. This protocol was
performed for Tf = 0.5 s yielding a total strain of γ = ∆γTf /trelax = 0.5.
We measured the elastic behavior of the network for each crosslinker stiffness by calculating
15
w, the strain energy density at each timestep:
(cid:32)(cid:88)
f
(cid:33)
Uxl
,
(cid:88)
xl
Uf +
(12)
w(t) =
1
XY
where Uf is the mechanical energy of individual filaments (Eq. 1) and Uxl is the mechanical energy
of each crosslink (Eq. 2). By averaging over windows of size trelax, we determine w(γ). Fig. 5
shows the results of these calculations for various values of kxl. For extremely low kxl, the strain
energy scaled linearly with strain, w ∝ γ, indicating that the network showed no resistance to
shear: G = d2w/dγ2 = 0. For high kxl, we observe a neo-Hookean strain stiffening behavior,
w ∝ γ4 (54). Thus, we can tune the material properties of crosslinked semiflexible networks from
being liquid-like, with w ∝ γ, through the Hookean elastic regime of w ∝ γ2 up to the strain
stiffening regimes of w ∝ γ3 and w ∝ γ3.5, as previously reported in experiments (47, 49). We
show in Section S4 and Fig. S5 that the monotonic increase in scaling for low kxl corresponds to
a regime where the strain energy is mostly stored in the crosslinkers, while the plateau at high kxl
corresponds to a regime where the strain energy is mostly stored in filaments.
Figure 5: Tunable elasticity of crosslinked networks. (A) Snapshots of a strained network (ka = 1000
pN/µm, kxl = 20pN/µm) at γ = 0.1, γ = 0.25, and γ = 0.5. Color indicates stretching energy on each
link, with green being the lowest and yellow being the highest. For all snapshots, t = γ × 1 s. (B) The
potential energy of the network as a function of time shown at different strains γ0 = 0.1 (circles), γ0 = 0.25
(squares), and γ0 = 0.4 (triangles), where t0 = γ0 × 1 s. Black dashed line shows the strain protocol.
(C) Strain energy density (w = U/XY ) for various values of crosslinker stiffness kxl. Blue dashed line
indicates expected behavior for a linearly elastic solid (w ∝ γ2) and green dashed line indicates strain
stiffening behavior of w ∝ γ3.5 as expected for semiflexible polymer networks (47, 50). (D) The power-law
exponent of w(γ) as a function of crosslinker stiffness, evaluated by least squares fitting ln (w) as a function
of ln (γ).
16
ABCDγ=0.5γ=0.25γ=0.14.3 Ensembles of motors interacting with individual filaments simulate
actin motility assays
While the attachment, detachment and speed of an individual myosin motor is a model input (de-
scribed in Crosslinkers and Motors, above), the collective action of many motors on a filament is
an output that can be compared with actin motility assays (36, 55). In the canonical motility assay
experiments, a layer of myosin is attached to a glass coverslip, and actin filaments are distributed
on top of the layer of myosin motors. The fixed motors translocate the actin filaments. The speed
of an actin filament has been reported to depend nonlinearly on the concentration of myosin and
the concentration of ATP in the sample (32, 33). Thus, by allowing the filaments to interact with
more motors, one can monotonically increase the filament speed to a constant value.
To explore the dynamics of such an assay, we randomly distributed motors on a 50 µm ×50 µm
periodic simulation cell and tethered one head of each motor to its initial position. These model
motors represent myosin minifilaments with dozens of heads, and therefore have a high default
duty ratio (rD = 0.5), and rest length lm = 0.5 µm (1, 56). Filaments were then introduced in the
simulation cell and allowed to interact with the free motor heads. The strength of motor-filament
interactions was manipulated in three ways: by varying the motor concentration ρm, the filament
m ). While L and rD are difficult to
contour length L, and the duty ratio rD = kon
modulate experimentally in a well-controlled fashion, as they require the addition of other actin-
binding proteins to the assay, they are predicted to impact the dynamics of actin by varying the
number of myosin heads bound to an actin filament at any one time (33). Since they are both
simple functions of the model's parameters, we were able to test this hypothesis directly. We plot
our simulation results in Fig. 6 as functions of the dimensionless control parameter M = ρmlmLrD
(where ρmlm is the linear motor density), which represents the average number of bound motor
heads per filament.
m + kof f
m /(kon
Our findings are qualitatively similar to the previously reported experimental results and ex-
pand on them by collapsing the trends observed while varying ρm, L, and rD into a single effective
parameter. At low M, i.e., low motor density, filament length, or duty ratio, Fig. 6B shows that
transverse motion dominates over longitudinal motion as the filament is not propelled by motors
faster than diffusion, and transverse filament fluctuations are larger than longitudinal fluctuations
(consistent with Fig. 3C). However, as M increases, longitudinal motion dominates. Consistent
with experimental results (31, 33), the longitudinal speed of the filament plateaus at v ≈ 1 µm/s,
which is the input unloaded speed of a single motor. In Fig. 6C, we plot the mean squared dis-
placement (MSD) of the filament, (cid:104)∆r2(cid:105) = (cid:104)(cid:126)r(t + δt) − (cid:126)r(t)2(cid:105) with angle brackets indicating an
average over time t. We show that low M yields diffusive behavior with (cid:104)∆r2(cid:105) ∝ δt, and high
M yields ballistic motion with (cid:104)∆r2(cid:105) ∝ δt2. We obtain similar results for motility assay behavior
with a corresponding Cytosim simulation, as shown in Section S5 (Fig. S6).
An interesting outcome of these simulations is how the direction of a filament changes over
time for varying M. Specifically, we calculate the directional autocorrelation of a filament and, in
turn, the persistence length of the path of the filament by applying Eq. 9 to the center of mass of
the filament at frames separated by ∆t = 1 s (Fig. 6D). Scaling arguments suggest that the path's
persistence length depends strongly on motor density, duty ratio, and filament length (57). In the
limit of high M, the distance between motors that are bound to a filament is sufficiently short
that the filament does not diffuse transversely; however, fluctuations in the filament configuration
still allow directional decorrelation, and consequently the path's persistence length is Lp. At low
17
M, the distance between bound motors is sufficiently large that rotational diffusion causes the
filament's path to be completely decorrelated, such that the path's persistence length approaches
0. In Fig. 6D, we show that the simulation agrees with theoretically predicted scaling laws at low
and high M (57). Our results delineate the values of M at which there are crossovers between the
predicted limiting regimes.
Figure 6: Nonlinear dependence of filament motility on motor-filament interaction probability. (A) Tra-
jectory of a filament for ρm = 4 µm−2 and L = 15 µm as a function of time for different values of the
duty ratio, rD. Depth of color indicates time of the snapshot, as indicated by the scale. Blue dot marks the
barbed end. (B) Filament speed decomposed into longitudinal (filled circles) and transverse (empty squares)
components as a function of the dimensionless parameter M = ρmlmLrD, by independently varying ρ
(green), L (red), and rD (blue). The default parameters were ρ = 4 µm−2, L = 15 µm, and rD = 0.5. (C)
Mean squared displacement for various values of M. Blue dashed line shows diffusive behavior and orange
dashed line shows ballistic behavior. (D) Path persistence length for simulations described in (B), evaluated
via Eq. 9 over 5 replicates. Dashed lines are theoretical predictions for these values using equations (1)-(6)
in (57).
4.4 Molecular motors cause flexible, crosslinked networks to contract
When motors, crosslinkers, and filaments are combined into a single assembly, simulated net-
works contract. The structure and dynamics of these networks exhibits a rich dependence on
motor and crosslinker densities, binding/unbinding kinetics, and stiffness parameters. Here, we
show one illustrative example to demonstrate that our model reproduces actomyosin contractility
for a reasonable choice of parameters (Fig. 7A). The network is initialized by randomly orient-
ing 500 filaments, each 10 µm long, within a 50 µm × 50 µm simulation cell. We distribute
0.15 µm long crosslinkers throughout the simulation cell at a density of 1 µm−2, and 0.5 µm
long motor oligomers at a density of 0.2 µm−2. As the simulation evolves, the actin density be-
18
BCDAtime (s)0 100 200rD=0.001rD=0.05rD=0.33rD=0.670.00.20.94.725.0comes more heterogenous as motors condense actin filaments into dense disordered aggregates.
This density heterogeneity can be quantified by the radial distribution function of actin filaments,
g(r) = P (r)/(2πrδrρf ), where P (r) is the probability that two filaments are separated by a dis-
tance r, δr = 0.1 µm is the spatial bin size, and ρf is the filament density. As shown by Fig. 7B,
g(r) ≈ 1 at t = 0 for all r as the actin filaments are homogeneously distributed. However, over
time it becomes more peaked at lower separation distances between filaments, indicating filament
aggregation.
To measure the contractile activity of the network, we evaluate the divergence of its velocity
field. This is done by calculating the velocity of each of the actin beads, followed by a grid-based
interpolation of a velocity vector field from those values (black arrows in Fig. 7C; interpolation
scheme described in Section S6). One can then evaluate the divergence ∇ · (cid:126)v of the interpolated
field at every spatial location (color of Fig. 7C). Since there is no flux of actin into the simulation
box, the total divergence of the flow field is zero at all times (i.e.,(cid:82) (∇ · (cid:126)v) dA = 0). Therefore, we
weight the divergence of each patch of the network by its local density and measure(cid:82) ρa(cid:104)∇ · (cid:126)v(cid:105)dA
where ρa = na/dA is the number density of actin beads and (cid:104)∇· (cid:126)v(cid:105) is the average actin divergence
in the patch of size dA. As shown in Section S6, this order parameter shows consistent behavior
for small patches (dA ≤ (10 µm)2) and a range of step sizes (h ≤ 20 s) for the velocity calculation
(Eq. S6). We also measure the average filament strain ∆s, in the network, where
(cid:33)
(cid:80)N
(cid:126)rN − (cid:126)r0
i=1 (cid:126)ri − (cid:126)ri−1
(cid:32)
∆s =
1 −
,
(13)
(cid:126)ri is the position of the ith bead on an (N + 1)-bead filament, and the bar denotes an average
over all filaments. Fig. 7D shows the results of measuring network divergence and filament strain
from 20 simulations with the same parameter choices as in Fig. 7A, but with different random
number seeds. The divergence measurement (blue) shows that the network is contractile, since
the density weighted divergence is negative, and its shape echoes the experimental results in (58),
where the magnitude of contractility decreases to a minimum before plateauing. The filament strain
measurement (red) shows that as the network is contracting, individual filaments are buckling. This
supports the notion that the mechanism behind contractility in disordered actomyosin networks is
actin filament buckling (59, 60). We note that, while the parameterization of motors that we used
for the motility assays yields contractile networks (Fig. S7), using a lower value of kof f
m resulted in
kinetics closer to those observed in experiment (58). This improvement with higher motor affinity
may reflect differences in the number of participating motor heads in contractility and motility
assays.
19
Figure 7: Contractility of a crosslinked filament network driven by motors. Filaments are red, motors
are black, and crosslinkers are green. (A) Network configurations at t = 0, 50, 150, and 398 s. While
all filaments are shown, only 10% of crosslinkers and 50% of motors are shown for clarity. (B) Radial
distribution function at frames corresponding to (A). (C) Quantification of the motion at t = 50 s. Arrows
(directions and sizes) indicate the filament-bead velocity field generated by the procedure in Section S6.
Colors map the corresponding divergence. (D) The density weighted divergence (blue; dA = (1 µm)2) and
average filament strain (red) of actin filaments for contractile networks. Dark lines for both curves shows
√
the mean µ(t) of these results at each time t over N = 20 simulations. Shaded areas show the standard error
of the mean µ(t) ± σ(t)/
N where σ(t) is the standard deviation.
5 Conclusion
In this paper, we have introduced an agent-based modeling framework that can accurately and
efficiently simulate active networks of filaments, motors, and crosslinkers to aid in the interpreta-
tion and design of experiments on cytoskeletal materials and synthetic analogs. While our focus
here has been on selecting parameters that are representative of the actin cytoskeleton, we expect
that this framework can be adapted to treating other active polymer assemblies as well, such as
microtubule-kinesin-dynein networks. We demonstrated that the model gives rise to both qual-
itative and quantitative trends for structure and dynamics observed in experiments and provides
experimentally testable predictions. Specifically, we reproduced the experimentally observed and
theoretically described fluctuation statistics of actin filaments. We also captured strain stiffening
scalings and predicted how network elasticity can potentially be tuned via crosslinker stiffness.
We modeled sliding filament assays and determined specific system parameters that lead to the
crossover from transversely diffusive to longitudinally processive motion first predicted in (57). In
separate studies, we use our model to explore the phase space of various network structures and
the dynamics that lead to them (3).
While our model captures many experimental observations, we simplified certain features to
20
0s50s150s398sABCDt=50slimit both computational cost and model complexity. First, the structure of myosin minifilaments
is significantly more complex than a two-headed spring. As mentioned, minifilaments have dozens
of heads, which allows them to attach to more than two filaments simultaneously, significantly
increasing local network elasticity (61) and enabling more complex motor dynamics (62). Second,
filaments do not polymerize, depolymerize, or sever in the simulations; it is clear, however, that
recycling of actin monomers, actin treadmilling and, to a lesser degree, filament severing play im-
portant roles in contraction and shape formation (60, 63). Third, our simulations are restricted to
2D, without steric or hydrodynamic interactions. This can play a role in motility assays, for exam-
ple, where at high actin densities, actin filaments organize into polar patterns with characteristic
autocorrelation times (64). It would be valuable to make the model a progressively more faithful
representation of reality in the future to better understand how each of these choices impacts the
behavior of the model and in turn the implications for the associated physics.
6 Author Contributions
S.L.F., S.B., G.M.H., and A.R.D. designed the research. S.L.F. designed and implemented the
software, and executed the calculations. S.B. provided a prototype program. S.L.F., S.B., G.M.H.,
and A.R.D. wrote the paper.
7 Acknowledgements
We thank M. Gardel, J. Weare, C. Matthews, E. Thiede, F. Nedelec, F.C. Mackintosh, and M.
Murrell for helpful conversations. We thank C. Tung, J. Harder, and S. Mallory for critical readings
of the manuscript. This research was supported in part by the University of Chicago Materials
Research Science and Engineering Center (NSF Grant No. 1420709). S.L.F. was supported by
the DoD through the NDSEG Program. G.M.H. was supported by an NIH Ruth L. Kirschstein
NRSA award (1F32GM113415-01). S.B. acknowledges support from the Institute for the Physics
of Living Systems at the University College London.
8 Supporting Citations
Reference (65) appears in the Supporting Material.
References
[1] Niederman, R., and T. D. Pollard, 1975. Human platelet Myosin II In vitro assembly and
structure of myosin filaments. The Journal of Cell Biology 67:72–92.
[2] Huxley, H., 1969. The mechanism of muscular contraction. Science 164:1356–1366.
[3] Stam, S., S. L. Freedman, S. Banerjee, K. L. Weirich, A. R. Dinner, and M. L. Gardel, 2017.
Filament Rigidity And Connectivity Tune The Deformation Modes Of Active Biopolymer Net-
works. bioRxiv http://biorxiv.org/content/early/2017/05/25/141796.
21
[4] MacKintosh, F., J. Kas, and P. Janmey, 1995. Elasticity of semiflexible biopolymer networks.
Physical Review Letters 75:4425.
[5] Head, D. A., A. J. Levine, and F. C. MacKintosh, 2003. Distinct regimes of elastic response and
deformation modes of cross-linked cytoskeletal and semiflexible polymer networks. Physical
Review E 68:061907.
[6] Wilhelm, J., and E. Frey, 2003. Elasticity of stiff polymer networks. Physical Review Letters
91:108103.
[7] Kim, T., W. Hwang, H. Lee, and R. D. Kamm, 2009. Computational analysis of viscoelastic
properties of crosslinked actin networks. PLoS Computational Biology 5:e1000439.
[8] Dasanayake, N. L., P. J. Michalski, and A. E. Carlsson, 2011. General mechanism of acto-
myosin contractility. Physical Review Letters 107:118101.
[9] Erdmann, T., and U. S. Schwarz, 2012. Stochastic force generation by small ensembles of
Myosin II motors. Physical Review Letters 108:188101.
[10] Wang, S., and P. G. Wolynes, 2012. Active contractility in actomyosin networks. Proceedings
of the National Academy of Sciences 109:6446–6451.
[11] Nedelec, F., and D. Foethke, 2007. Collective Langevin dynamics of flexible cytoskeletal
fibers. New Journal of Physics 9:427.
[12] Ennomani, H., G. Letort, C. Gu´erin, J.-L. Martiel, W. Cao, F. N´ed´elec, M. Enrique, M. Th´ery,
and L. Blanchoin, 2016. Architecture and Connectivity Govern Actin Network Contractility.
Current Biology 26:616–626.
[13] Gordon, D., A. Bernheim-Groswasser, C. Keasar, and O. Farago, 2012. Hierarchical self-
organization of cytoskeletal active networks. Physical Biology 9:026005.
[14] Kim, T., 2014. Determinants of contractile forces generated in disorganized actomyosin
bundles. Biomechanics and Modeling in Mechanobiology 14:345–355.
[15] Popov, K., J. Komianos, and G. A. Papoian, 2016. MEDYAN: Mechanochemical Simula-
tions of Contraction and Polarity Alignment in Actomyosin Networks. PLoS Computational
Biology 12:e1004877.
[16] Stam, S., J. Alberts, M. L. Gardel, and E. Munro, 2015. Isoforms: Confer Characteristic Force
Generation and Mechanosensation by Myosin II Filaments. Biophysical Journal 108:1997–
2006.
[17] Cyron, C., K. Muller, K. Schmoller, A. Bausch, W. Wall, and R. Bruinsma, 2013. Equi-
librium phase diagram of semi-flexible polymer networks with linkers. Europhysics Letters
102:38003.
[18] Muller, K. W., C. J. Cyron, and W. A. Wall, 2015. Computational analysis of morphologies
and phase transitions of cross-linked, semi-flexible polymer networks. Proceedings of the
Royal Society A 471:20150332.
22
[19] Allen, M. P., and D. J. Tildesley, 1989. Computer Simulation of Liquids. Oxford University
Press.
[20] Plimpton, S., 1995. Fast parallel algorithms for short-range molecular dynamics. Journal of
Computational Physics 117:1–19.
[21] Rubinstein, M., and R. H. Colby, 2003. Polymer Physics. OUP Oxford.
[22] Ott, A., M. Magnasco, A. Simon, and A. Libchaber, 1993. Measurement of the persistence
length of polymerized actin using fluorescence microscopy. Physical Review E 48:R1642.
[23] Kojima, H., A. Ishijima, and T. Yanagida, 1994. Direct measurement of stiffness of single
actin filaments with and without tropomyosin by in vitro nanomanipulation. Proceedings of
the National Academy of Sciences 91:12962–12966.
[24] Higuchi, H., T. Yanagida, and Y. E. Goldman, 1995. Compliance of thin filaments in skinned
fibers of rabbit skeletal muscle. Biophysical Journal 69:1000.
[25] Leimkuhler, B., and C. Matthews, 2015. Molecular Dynamics: with Deterministic and
Stochastic Numerical Methods. Springer.
[26] Vale, R. D., F. Malik, and D. Brown, 1992. Directional instability of microtubule transport in
the presence of kinesin and dynein, two opposite polarity motor proteins. The Journal of Cell
Biology 119:1589–1596.
[27] Bourdieu, L., T. Duke, M. Elowitz, D. Winkelmann, S. Leibler, and A. Libchaber, 1995.
Spiral defects in motility assays: a measure of motor protein force. Physical review letters
75:176.
[28] Metropolis, N., A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, and E. Teller, 1953.
Equation of state calculations by fast computing machines. The Journal of Chemical Physics
21:1087–1092.
[29] Gillespie, D. T., 1977. Exact stochastic simulation of coupled chemical reactions. The Journal
of Physical Chemistry 81:2340–2361.
[30] N´ed´elec, F., 2002. Computer simulations reveal motor properties generating stable antiparal-
lel microtubule interactions. The Journal of Cell Biology 158:1005–1015.
[31] Kron, S. J., and J. A. Spudich, 1986. Fluorescent actin filaments move on myosin fixed to a
glass surface. Proceedings of the National Academy of Sciences 83:6272–6276.
[32] Umemoto, S., and J. R. Sellers, 1990. Characterization of in vitro motility assays using
smooth muscle and cytoplasmic myosins. Journal of Biological Chemistry 265:14864–14869.
[33] Harris, D. E., and D. Warshaw, 1993. Smooth and skeletal muscle myosin both exhibit low
duty cycles at zero load in vitro. Journal of Biological Chemistry 268:14764–14768.
[34] Finer, J. T., R. M. Simmons, J. A. Spudich, et al., 1994. Single myosin molecule mechanics:
piconewton forces and nanometre steps. Nature 368:113–119.
23
[35] Roux, B., 2011. Molecular Machines. World Scientific.
[36] Riveline, D., A. Ott, F. Julicher, D. A. Winkelmann, O. Cardoso, J.-J. Lacap`ere,
S. Magn´usd´ottir, J.-L. Viovy, L. Gorre-Talini, and J. Prost, 1998. Acting on actin: the electric
motility assay. European Biophysics Journal 27:403–408.
[37] Leimkuhler, B., and C. Matthews, 2013. Robust and efficient configurational molecular sam-
pling via Langevin dynamics. The Journal of Chemical Physics 138:174102.
[38] Prost, J., 1995. The physics of liquid crystals. 83. Oxford university press.
[39] Odijk, T., 1986. Theory of lyotropic polymer liquid crystals. Macromolecules 19:2313–2329.
[40] Murrell, M., T. Thoresen, and M. Gardel, 2014. Reconstitution of contractile actomyosin
arrays. Methods in enzymology 540:265.
[41] Bird, R. B., R. C. Armstrong, and O. Hassager, 1987. Dynamics of polymeric liquids. Vol. 1:
Fluid mechanics .
[42] Veigel, C., J. E. Molloy, S. Schmitz, and J. Kendrick-Jones, 2003. Load-dependent kinetics
of force production by smooth muscle myosin measured with optical tweezers. Nature Cell
Biology 5:980–986.
[43] Ferrer, J. M., H. Lee, J. Chen, B. Pelz, F. Nakamura, R. D. Kamm, and M. J. Lang, 2008.
Measuring molecular rupture forces between single actin filaments and actin-binding proteins.
Proceedings of the National Academy of Sciences 105:9221–9226.
[44] Stricker, J., T. Falzone, and M. L. Gardel, 2010. Mechanics of the F-actin cytoskeleton.
Journal of Biomechanics 43:9–14.
[45] Frontali, C., E. Dore, A. Ferrauto, E. Gratton, A. Bettini, M. Pozzan, and E. Valdevit, 1979.
An absolute method for the determination of the persistence length of native DNA from elec-
tron micrographs. Biopolymers 18:1353–1373.
[46] Everaers, R., F. Julicher, A. Ajdari, and A. Maggs, 1999. Dynamic fluctuations of semiflexi-
ble filaments. Physical Review Letters 82:3717.
[47] Gardel, M., J. Shin, F. MacKintosh, L. Mahadevan, P. Matsudaira, and D. Weitz, 2004. Elastic
behavior of cross-linked and bundled actin networks. Science 304:1301–1305.
[48] Koenderink, G., M. Atakhorrami, F. MacKintosh, and C. Schmidt, 2006. High-frequency
stress relaxation in semiflexible polymer solutions and networks. Physical Review Letters
96:138307.
[49] Kasza, K., G. Koenderink, Y. Lin, C. Broedersz, W. Messner, F. Nakamura, T. Stossel,
F. MacKintosh, and D. Weitz, 2009. Nonlinear elasticity of stiff biopolymers connected by
flexible linkers. Physical Review E 79:041928.
24
[50] Lin, Y.-C., N. Y. Yao, C. P. Broedersz, H. Herrmann, F. C. MacKintosh, and D. A. Weitz,
2010. Origins of elasticity in intermediate filament networks. Physical Review Letters
104:058101.
[51] Bustamante, C., J. Marko, E. Siggia, and S. Smith, 1994. Entropic Elasticity of Slambda
S-Phage DNA. Science 265:1599.
[52] Marko, J. F., and E. D. Siggia, 1995. Stretching DNA. Macromolecules 28:8759–8770.
[53] Evans, D. J., and G. Morriss, 1984. Nonlinear-response theory for steady planar Couette flow.
Physical Review A. 30:1528.
[54] Shokef, Y., and S. A. Safran, 2012. Scaling laws for the response of nonlinear elastic media
with implications for cell mechanics. Physical Review Letters 108:178103.
[55] Walcott, S., D. M. Warshaw, and E. P. Debold, 2012. Mechanical coupling between
myosin molecules causes differences between ensemble and single-molecule measurements.
Biophysical Journal 103:501–510.
[56] Ideses, Y., A. Sonn-Segev, Y. Roichman, and A. Bernheim-Groswasser, 2013. Myosin II
does it all: assembly, remodeling, and disassembly of actin networks are governed by myosin
II activity. Soft Matter 9:7127–7137.
[57] Duke, T., T. E. Holy, and S. Leibler, 1995. "Gliding assays" for motor proteins: A theoretical
analysis. Physical Review Letters 74:330.
[58] Murrell, M., and M. L. Gardel, 2014. Actomyosin sliding is attenuated in contractile
biomimetic cortices. Molecular Biology of the Cell 25:1845–1853.
[59] Lenz, M., T. Thoresen, M. L. Gardel, and A. R. Dinner, 2012. Contractile units in disordered
actomyosin bundles arise from F-actin buckling. Physical Review Letters 108:238107.
[60] Murrell, M. P., and M. L. Gardel, 2012. F-actin buckling coordinates contractility and sev-
ering in a biomimetic actomyosin cortex. Proceedings of the National Academy of Sciences
109:20820–20825.
[61] Linsmeier, I., S. Banerjee, P. W. Oakes, W. Jung, T. Kim, and M. Murrell, 2016. Disordered
actomyosin networks are sufficient to produce cooperative and telescopic contractility. Nature
Communications 7:12615.
[62] Scholz, M., S. Burov, K. L. Weirich, B. J. Scholz, S. A. Tabei, M. L. Gardel, and A. R. Dinner,
2016. Cycling State that Can Lead to Glassy Dynamics in Intracellular Transport. Physical
Review X 6:011037.
[63] Wilson, C. A., M. A. Tsuchida, G. M. Allen, E. L. Barnhart, K. T. Applegate, P. T. Yam,
L. Ji, K. Keren, G. Danuser, and J. A. Theriot, 2010. Myosin II contributes to cell-scale actin
network treadmilling through network disassembly. Nature 465:373–377.
[64] Schaller, V., C. Weber, C. Semmrich, E. Frey, and A. R. Bausch, 2010. Polar patterns of
driven filaments. Nature 467:73–77.
25
[65] Hetland, R., and J. Travers, 2001. SciPy: Open source scientific tools for Python: rbf - Radial
basis functions for interpolation/smoothing scattered Nd data.
26
A versatile framework for simulating the dynamic
mechanical structure of cytoskeletal networks:
Supporting Material
S. L. Freedman, S. Banerjee, G. M. Hocky, A. R. Dinner
S1 Calculation of crosslinker head position during binding and
unbinding
In this section, we describe how we update the binding state (I1(2) in Eq. 2) and position ((cid:126)r1(2)) of
a crosslinker head. The binding states and positions of the two heads of a crosslinker are coupled
only through the potential energy (Eq. 2).
We first discuss binding. An unbound crosslinker head with position (cid:126)ru can attempt to bind to
the closest point on each nearby filament link. Let (cid:126)li = (cid:126)ri − (cid:126)ri−1, where (cid:126)ri is the position of the ith
bead on the filament to which the link belongs. Then, we propose a bound state with binding point
(cid:126)li = 0 or p ≤ 0
p ≥ 1
otherwise
(S1)
(cid:126)ri−1
(cid:126)rb =
(cid:126)ri
(cid:126)ri−1 + p(cid:126)li
where p = ((cid:126)ru − (cid:126)ri) · (cid:126)li. Eq. S1 can be interpreted easily in a reference frame in which (cid:126)li is
oriented vertically (Fig. S1A): if (cid:126)ru is below the link, (cid:126)rb = (cid:126)ri−1; if it is above the filament then
(cid:126)rb = (cid:126)ri; otherwise (cid:126)rb is the intersection of (cid:126)li with the line perpendicular to (cid:126)li that passes through
If (cid:126)rb − (cid:126)ru < rc, the changes in binding state and position are accepted with probability
(cid:126)ru.
xl ∆t)P of f→on
(kon
(see main text, Crosslinkers).
xl,i
1
Figure S1: Position of crosslinker head upon binding or unbinding. (A) Any crosslinker head in the aqua,
yellow, and gray areas (such as the filled blue, green, and black circles) can bind to the blue, green, and
black binding points (circles with crosses), respectively. (B) The process by which a crosslinker generates
an unbinding point (ru) at time t + h using its original displacement at time t when it snapped to the binding
point rb.
For unbinding, we do the following. At the time of binding (t), we record the displacement
vector, (cid:126)rbu = (cid:126)rb(t) − (cid:126)ru(t), and the vector connecting the ends of the filament link, (cid:126)li(t) = (cid:126)ri(t) −
(cid:126)ri−1(t). At the time that we attempt unbinding (t + h), we determine the angle of rotation of the
filament link:
(cid:126)li(t)(cid:126)li(t + h)
Then, the position to which the crosslinker head tries to jump is
θ = arccos
(cid:126)ru(t + h) = (cid:126)rb(t + h) −
(cid:32) (cid:126)li(t) · (cid:126)li(t + h)
(cid:33)
(cid:18)cos (θ) − sin (θ)
.
sin (θ)
cos (θ)
(S2)
(S3)
(cid:126)rbu(t)
(cid:19)
as shown in Fig. S1B. This jump is accepted with probability (kof f
. The motivation
for this scheme is that it ensures that a head that jumps onto (off) a filament link returns to its
original position if it unbinds (rebinds) immediately. Detailed balanced consistent with Eq. 2 can
thus be satisfied through the acceptance probabilities (kon
and (kof f
xl ∆t)P on→of f
xl ∆t)P of f→on
xl,i
xl,i
.
xl ∆t)P on→of f
xl,i
S2 Relaxation times scales
In this section, we present data on filament and network time scales that inform our choices of
sampling frequencies.
S2.1 Decorrelation of filament angles
The evaluations of persistence length in Actin filaments exhibit predicted spatial and temporal
fluctuations in the main text average over independent configurations of filaments. To determine
the amount of time between independent configurations in a trajectory of a single filament, we
evaluated the integrated autocorrelation time of the angles θi for i ∈ [2 . . . 20] between links along
2
ABa 21 bead filament. Fig. S2A shows the autocorrelation
R(θ, s) =
(cid:104)θ(t)θ(t + s)(cid:105) − (cid:104)θ(t)(cid:105)2
(cid:104)θ(t)2(cid:105) − (cid:104)θ(t)(cid:105)2
(S4)
where s is the time between realizations and the angle brackets represent an average over all 19
angles and all 1900 saved configurations. Fig. S2B shows the integrated autocorrelation time τ as
a function of the simulation cutoff time tf inal, where
(cid:90) tf inal
τ (θ) =
R(θ, s)ds.
(S5)
For all choices of tf inal, τ < 2 s and therefore configurations that are separated by at least 2 s
should be independent realizations with respect to angles between subsequent filament links.
0
Figure S2: Estimation of the characteristic decorrelation time for persistence length measurements. (A)
Decorrelation of angles between filament links for a 21 bead filament with ka = 1 pN/µm, la = 1 µm, and
κB = 0.068 pNµm2. (B) Measurement of the integrated autocorrelation time τ for different values of the
cutoff time tf inal.
S2.2 Shear relaxation times
One extra parameter that must be set for shear simulations is the relaxation time (trelax)-i.e.,
the minimum time between strain steps for responses to be history independent. We probed this
question computationally by determining if the parameter of interest (total potential energy of
filaments and crosslinkers) varied significantly for different periods of relaxation between steps
of ∆γ = 0.001. Fig. S3 shows that while very small trelax values do yield higher energies at
equivalent strains, as trelax is increased, the curves collapse for identical strains.
In the shear
simulations in the main text (Tunable elastic behavior of crosslinked filament networks), trelax = 1
ms (yellow curve).
3
ABFigure S3: Total potential energy as a function of strain for various relaxation times. Simulation parameters,
are otherwise identical to the shear simulations in the main text.
S3 Comparison with CytoSim
Cyotsim is a freely available C++ software package developed to simulate active polymer networks
and described in (1). While AFiNeS shares many of the same features, for clarity we enumerate
the technical differences.
• The filament model. AFiNeS uses a bead spring chain and CytoSim uses a chain con-
strained via Lagrange multipliers.
• Attachment of motors and crosslinkers. CytoSim uses a continuous-time Monte Carlo
procedure (the Gillespie algorithm (2)) to calculate when a motor should attempt attach-
ment to a filament, while AFiNeS attempts with the probability computed for each discrete
timestep of fixed duration. In Cytosim, the attachment of a motor to a filament is not depen-
dent on the distance from the filament, other than that it must be below a threshold, whereas
in AFiNeS, a closer motor has a higher probability of attachment, due to detailed balance
considerations.
• Detachment of motors and crosslinkers. CytoSim has a force dependent detachment of
crosslinkers. This was not a necessary detail to reproduce the benchmarks shown in the re-
sults section, and detailed balance would require altering the motor and crosslinker dynam-
ics, so we have not included it in the present version. We plan in the future to understand
how this detail effects cytoskeletal networks in general and add it as an option to AFiNeS.
• Capabilities present in one and not the other. AFiNeS implements network shearing.
CytoSim implements filament polymerization and depolymerization, microtubule asters, and
spherical geometries.
To compare the two packages, we have used CytoSim to run the benchmarks associated with
filament fluctuations (Fig. S4) and motility assays (Fig. S6, below). For the filament fluctuation
benchmarks, shown in Fig. S4, we find that while CytoSim is able to yield nearly the correct
persistence length of filaments, at long segment lengths it performs worse than AFiNeS, perhaps
because it uses linearized versions of the angle forces (1).
4
Figure S4: Measurements of persistence length for CytoSim filaments (red) compared with the same mea-
surements for AFiNeS (blue). (A) Cosine correlation function and ∆θ2 correlation function for 20 CytoSim
fibers with Lp = 17 µm fluctuating for 2000 seconds. See Section 4.1 of the main text for details. (B)
Measurement of Lp as function of segment length, la, using the fit to the first 5 data points of (cid:104)∆θ2(cid:105) in (A).
(C) Measurement of Lp as a function of input bending modulus for CytoSim and AFiNes. Colors are the
same as panel B.
S4 Parsing the energy in sheared networks
To further examine the source of the energy scalings shown in Fig. 5D, we measure the fraction
of the total energy density w from each of its sources in the network, the stretching energy of
filaments, the stretching energy of crosslinkers, and the bending energy of filaments, as shown
in Fig. S5.
In general, we find that shearing the network stretches and bends actin filaments,
and also stretches crosslinkers, as in Fig. S5A-C. Fig. S5B-C show that, as crosslinkers become
more stiff, more of the energy from the strain is concentrated on the filaments. Fig. S5A shows
that for crosslinkers, the trend is not monotonic. When kxl < 100 pN/µm, increasing crosslinker
stiffness results in more energy in the crosslinkers, and in this regime, the scaling of w(γ) increases
monotonically. However, for kxl ≥ 100 pN/µm, the trend reverses, and the strain energy density
concentrates on the filaments more than the crosslinkers, as seen in Fig. S5D-E. In this regime,
the scaling of w(γ) plateaus near the value 3.5, reflecting the prediction for the differential shear
modulus in a strain controlled rheology experiment, G = d2w/dγ2 ∝ γ3/2 (3).
5
ABCFigure S5: Absolute (A-C) and relative (D-F) energy contributions from crosslinkers stretching (A, D),
filaments stretching (B, E), and filaments bending (C, F) for the sheared network discussed in Tunable
elastic behavior of crosslinked filament networks.
S5 Comparison with Cytosim for motility assays
We also used Cytosim to simulate the motility assays described in the main text (Ensembles of
motors interacting with individual filaments simulate actin motility assays). The results, shown in
Fig. S6, are generally congruent with the results from AFiNeS in Fig. 6. We find that increasing
motor density, filament length, and duty ratio increase longitudinal motion and decrease transverse
motion of the filament (Fig. S6B), and makes the filament move more ballistically (Fig. S6C).
Furthermore, the path persistence length plots (Fig. S6D) are nearly identical to the measurements
obtained using AFiNES. Thus, it is reassuring that the two models agree to this extent despite the
differences in filament and binding implementations.
6
ABCFEDFigure S6: Motility measurements at varying motor density, filament length, and duty ratio generated using
CytoSim. For a detailed description of this calculation see main text, Ensembles of motors interacting with
individual filaments simulate actin motility assays.
S6 Procedure for quantifying contractility
An actin assay can be considered contractile if it has regions to which most of the actin aggregates.
In an experiment with a limited field of view, the net flux of actin into the field of view is positive
when the system is contractile. This flux corresponds mathematically to a negative value for the
integral of the divergence of the velocity field over the area (4, 5). However, in our simulations,
all particles' positions are known and there is no flux of material into or out of the simulation
region owing to the periodic boundary condition. Thus the total divergence obtained by integrating
over the simulation box must be zero. Nevertheless, we can still compute the density-weighted
divergence to quantify contractility, as we now describe.
To ensure that the divergence is well-defined at all points, we first interpolate a continuous ve-
locity field. When the data are experimental images, the velocity field is determined using Particle
Image Velocimetry (PIV). Here, we take a similar approach, with the advantage that positions of
actin beads are a direct output of the simulation, analogous to tracer particles in experiments. To
this end, for each filament bead i with position (cid:126)ri(t) at time t, we calculate the velocity by forward
finite difference:
(cid:126)vi((cid:126)ri, t) =
,
(S6)
(cid:126)ri(t + h) − (cid:126)ri(t)
h
where h is a suitable amount of time to characterize motion. We calculate the average velocity of
each (5 µm)2 bin. Similarly to PIV, we lower the noise further by setting a threshold, and only
consider bins with at least n actin beads. We then interpolate the bin values with Gaussian radial
basis functions (RBFs):
(cid:126)v((cid:126)r) =
(cid:126)wke−((cid:126)r−(cid:126)rk/)2
(S7)
M(cid:88)
k=1
7
BCDAtime (s)0 100 200rD=0.001rD=0.05rD=0.33rD=0.67where M is the number of bins with at least n actin beads, is a constant related to the width of the
Gaussian RBFs, and (cid:126)wk are their weights. The optimal value for is generally close to the value
of the average distance between RBFs (6); we found = 5 µm and a threshold of n = 10 yielded
a robust interpolation across many different actin structures. We use the scipy.interpolate.Rbf
Python package to determine the weights (6). We calculate the divergence of the resulting field
dvx((cid:126)r)/dx + dvy((cid:126)r)/dy by using finite difference approximations for the derivatives of Eq. S7.
Examples of this velocity field and the local divergence are shown in Fig. 7C and Fig. S7C.
As noted above, given ∇ · (cid:126)v, we quantify the contractility by the density weighted divergence,
(cid:82) ρa(cid:104)∇ · (cid:126)v(cid:105)dA. In Fig. S7E we show an example where the density weighting has the effect of
significantly increasing the magnitude of the areas with negative divergence. To understand how
the contractility varies with length scale, we replace the integral with the sum over square regions
(cid:88)
ρa((cid:126)rk)(cid:104)∇ · (cid:126)v(cid:105)kdA
(S8)
k
and vary the size of the regions, dA = dxdy (Fig. S7F). For the maximum size dA = (50 µm)2
(yellow curve), the density weighted divergence fluctuates around 0 as expected from the zero actin
flux. However for region sizes dA ≤ (10 µm)2, the values are consistently negative, indicating
contractility; the curves decrease to a minimum before plateauing closer to 0, as seen in experiment
(5). We also show, in Fig. S7G, that the trend of this order parameter is independent of the time
scale h used to calculate the velocity in Eq. S6.
8
Figure S7: Calculation of density weighted divergence for a simulated contractile actomyosin network. (A-
m = 1 s−1, and ρm = 1 µm−2. In (A), all filaments and
D) Identical to Fig. 7, but with kof f
10% of motors and crosslinkers are shown. (E) Same as (C), but the color is weighted by the actin density
ρa. (F) Dependence of the density weighted divergence on the patch size used for integration, dA = dxdy,
with h = 10 s. (G) Dependence of the density weighted divergence on the time scale h used in calculating
the velocity of actin v in Eq. S6 with dx = dy = 1 µm.
m = 10 s−1, kend
Supporting References
[1] Nedelec, F., and D. Foethke, 2007. Collective Langevin dynamics of flexible cytoskeletal
fibers. New Journal of Physics 9:427.
[2] Gillespie, D. T., 1977. Exact stochastic simulation of coupled chemical reactions. The Journal
of Physical Chemistry 81:2340–2361.
[3] Gardel, M., J. Shin, F. MacKintosh, L. Mahadevan, P. Matsudaira, and D. Weitz, 2004. Elastic
behavior of cross-linked and bundled actin networks. Science 304:1301–1305.
9
0s100s250s400sABCDt=250smHzEFG[4] Murrell, M. P., and M. L. Gardel, 2012. F-actin buckling coordinates contractility and sev-
ering in a biomimetic actomyosin cortex. Proceedings of the National Academy of Sciences
109:20820–20825.
[5] Murrell, M., and M. L. Gardel, 2014. Actomyosin sliding is attenuated in contractile
biomimetic cortices. Molecular Biology of the Cell 25:1845–1853.
[6] Hetland, R., and J. Travers, 2001. SciPy: Open source scientific tools for Python: rbf - Radial
basis functions for interpolation/smoothing scattered Nd data.
10
|
1608.00363 | 2 | 1608 | 2017-04-07T16:03:53 | Curvature-guided motility of microalgae in geometric confinement | [
"physics.bio-ph",
"cond-mat.soft"
] | Microorganisms, such as bacteria and microalgae, often live in habitats consisting of a liquid phase and a plethora of interfaces. The precise ways in which these motile microbes behave in their confined environment remain unclear. Using experiments, Brownian dynamics simulations, and analytical theory, we study the motility of a single Chlamydomonas microalga in an isolated microhabitat with controlled geometric properties. We demonstrate how the geometry of the habitat controls the cell's navigation in confinement. The probability of finding the cell swimming near the boundary scales linearly with the wall curvature, as seen for both circular and elliptical chambers. The theory, utilizing an asymmetric dumbbell model of the cell and steric wall interactions, captures this curvature-guided navigation quantitatively with no free parameters. | physics.bio-ph | physics |
Curvature-guided motility of microalgae in geometric confinement
Tanya Ostapenko,1 Fabian Jan Schwarzendahl,1 Thomas Boddeker,1 Christian
Titus Kreis,1 Jan Cammann,1 Marco G. Mazza,1 and Oliver Baumchen1, ∗
1Max Planck Institute for Dynamics and Self-Organization (MPIDS), Am Fassberg 17, D-37077 Gottingen, Germany
Microorganisms, such as bacteria and microalgae, often live in habitats consisting of
a liquid phase and a plethora of interfaces. The precise ways in which these motile
microbes behave in their confined environment remain unclear. Using experiments,
Brownian dynamics simulations, and analytical theory, we study the motility of a single
Chlamydomonas microalga in an isolated microhabitat with controlled geometric prop-
erties. We demonstrate how the geometry of the habitat controls the cell's navigation
in confinement. The probability of finding the cell swimming near the boundary scales
linearly with the wall curvature, as seen for both circular and elliptical chambers. The
theory, utilizing an asymmetric dumbbell model of the cell and steric wall interactions,
captures this curvature-guided navigation quantitatively with no free parameters.
INTRODUCTION
Life in complex geometries manifests itself at the mi-
croscopic level through the myriad of ways in which mi-
croorganisms interact with their environment. This en-
tails a broad spectrum of microbiological phenomena,
ranging from amoebic crawling [1, 2] and fibroblast mi-
gration [3] guided by nano-patterned substrates, the di-
rectional migration of epithelial cells on curved surfaces
[4] and microbial proliferation in space-limited environ-
ments [5] to the motility of microbiological swimmers in
confinement [6]. In fact, the natural habitats for micro-
bial life are often non-bulk situations, including aqueous
microdroplets [7] and the interstitial space of porous me-
dia, such as rocks [8, 9] and soil [10]. The study of how
self-propelled microorganisms in a liquid medium inter-
act with their geometric boundaries finds application in
physiology with regards to spermatozoa motility in the
reproductive tract [11–14], the motion of parasites in the
vertebrate bloodstream [15], and in microbiology in the
context of biofilm formation [16–19].
This lends itself to inquiry over the mechanisms
involved in the interplay between the microorganism
and its confining domain. While walkers and crawlers
achieve locomotion by momentum transfer to a sub-
strate, i.e. friction, microorganisms swimming in fluid
medium might undergo long-range hydrodynamic inter-
actions, in addition to contact interactions [20, 21]. For
these microswimmers, a distinction between "puller"-
and "pusher"-type swimmers is required [22], since the
flow fields around the two classes entail fundamental dif-
ferences [23–26]. Though much work was reported on the
onset of collective effects of microswimmers due to reori-
entation, or lack thereof, at an interface (e.g. [27, 28]),
there are few studies on how a single microswimmer cell
interacts with a boundary in the absence of cell-cell in-
teractions. At flat interfaces, the contact of a spermato-
zoon's flagellum with a surface tends to rotate it towards
a boundary, thus preventing its escape from flat or weakly
curved surfaces [29]. However, for the puller-type mi-
croswimmer Chlamydomonas, a microalga with two an-
terior flagella, steric interactions and multiple flagellar
contacts were found responsible for its microscopic scat-
tering off of a flat interface [29]. Single scattering events
of an individual Chlamydomonas cell were also reported
at convex interfaces, where two regimes emerge as the
cell scatters off: an initial, contact force regime and a
second, hydrodynamics-dominated regime [30]. Beyond
these details of the microscopic interactions at interfaces,
the way in which the global swimming behavior of a sin-
gle cell is affected by the geometry of a confining domain
remains elusive.
In this article, we report on the motility of a single
Chlamydomonas reinhardtii cell in tailor-made microhab-
itats to elucidate the effects of geometric confinement.
We find that the dominant attributes of the swimming
statistics are the alga's spatial confinement, which lim-
its its motion to its swimming plane, and the compart-
ment's curved boundary in this plane. We study pre-
cisely a single isolated cell in order to exclude any cell-
cell interactions or collective effects. Our experiments are
in quantitative agreement with Brownian dynamics sim-
ulations and analytical theory, whose main ingredients
are steric wall interactions and the alga's torque at the
compartment interface during a finite interaction time.
While a conclusive description of the microscopic details
of wall interactions might remain debated today, our re-
sults illuminate how a single puller-type cell's navigation
in confinement is primarily dominated by the details of
the environment's geometric constraints.
RESULTS
We employed optical microscopy techniques to study
the motility of an individual Chlamydomonas cell con-
tained within an isolated quasi-two-dimensional microflu-
idic compartment. Experiments were performed in circu-
lar compartments with radii ranging from 25 to 500 µm,
2
The mean-squared displacement (MSD) represents a
parameter through which the type of swimming behav-
ior that the alga exhibits is characterized. Here, the
MSD for the observation time t was extracted from a
single alga's experimental trajectory for each compart-
ment size; see Fig. 1C. An immediate observation is that
the MSD curves show no clear transition between bal-
listic behavior, i.e. MSD ∼ t2, on short time scales to
diffusive, i.e. MSD ∼ t, on long time scales, as reported
in previous studies on Chlamydomonas swimming in un-
confined 2D environments (transition time from ballistic
to diffusive ∼ 2 s) [29]. A linear fit to the initial regime of
the experimental data yields an exponent of 1.90±0.03,
in agreement with a regime of ballistic swimming. The
exponent for rc = 25 µm of 1.7±0.03 is attributed to
the strong confinement situation, as the compartment
is not more than approximately 2 times larger than the
alga (body and flagella), representing the lower limit of
our statistical analysis. On long time scales, the MSD
reaches a plateau corresponding to the explorable area of
its confined environment. Hence, we find that the alga's
run-and-tumble-like motion in environments unconfined
in the swimming plane [31] becomes predominantly bal-
listic swimming in confinement.
The experimental cell trajectories were statistically av-
eraged and converted into relative probability density
maps for each compartment size. Figure 2 displays
a series of two-dimensional heat maps of the relative
probability density of the cell's positions for different
compartment sizes (see Materials and Methods for de-
tails). Our experimental data provides evidence for a pro-
nounced near-wall swimming effect inside the compart-
ment, whose significance decreases for increasing com-
partment size.
This near-wall swimming effect is further quantified
by azimuthally collapsing the heat maps from Fig. 2 into
radial probability densities, P (r), as depicted in Fig. 3.
We define P (r) as:
P (r) =
h(r)/2πr∆r
h(r)
2πr∆r dr
R rc
0
(1)
we normalize P (r) such that R rc
where r is the distance from the center of the compart-
ment, and h(r) is the count of all the alga's positions in
a circular shell at distance r with thickness ∆r. In or-
der to compare data from different compartment sizes,
0 P (r)dr = 1. Note that
a homogeneous distribution of trajectory points would
result in P (r) = 1/rc = const. by this definition.
In
Fig. 3, particularly the inset, we observe that P (r) starts
from a plateau in proximity of the compartment's center
and increases significantly close to the wall. The lateral
extent (full-width-half-maximum) of the peak of P (r)
ranges from 3–5 µm, about half a cell body diameter; the
peak position is consistently 9–11 µm away from the wall.
At the compartment wall, P (r) drops off, representing a
20μm
(A)
(B)
r
r
r
c
c
c
= 25 µm
= 50µm
= 100µm
= 150µm
= 500µm
r
c
r
c
fit: ~t2
104
102
2
m
µ
/
D
S
M
100
100
101
102
Time t / s
(C)
103
FIG. 1. Experimental design and trajectory analysis.
(A)
Optical micrograph of a single alga contained in a quasi-2D
circular compartment. (B) Exemplar single-cell trajectory for
rc = 50 µm. (C) Mean-squared displacements (solid lines) for
different compartment radii. The dashed line is a ∼ t2 best
fit to the short-time ballistic behavior.
as well as elliptical chambers. The height of all compart-
ments was approximately 20 µm, about one cell diameter
(body and flagella); thus, out-of-plane reorientations of
the cell are inhibited. Each experiment with a single,
isolated cell was repeated up to 10 times using different
cells each time. Trajectories were extracted using image
processing and particle tracking algorithms (see Materi-
als and Methods). Each cell trajectory corresponds to
a single, independent experiment for a particular com-
partment size, with no cell-cell interactions or collective
effects. The large amount of experimental data provides
reliable statistics, which are quantitatively compared to
Brownian dynamics simulations and analytical theory.
Motility in Circular Compartments
Figure 1A displays an image from a single experiment
(see movie S1) for a compartment radius rc = 50 µm,
from which the trajectory of the alga's body center was
extracted (Fig. 1B). The alga's trajectory shows a higher
density of trajectory points closer to the concave inter-
face, as compared to the center of the compartment. We
note that this higher density of points corresponds to an
enhanced probability of finding the alga within the vicin-
ity of the wall, and examine this quantity in detail during
the course of our study.
3
FIG. 2. Relative probability density for a single cell in circular confinement. Heat maps represent the alga's position obtained
from experimental data for different compartment sizes: (L–R) rc = 25 µm, 50 µm, 100 µm, 150 µm, 500 µm. Each map contains
statistically averaged data from a minimum of 2–5 independent experiments. For the applied normalization, see Materials and
Methods.
r = 25 m
c
r = 50 m
c
r = 75 m
c
r = 100 m
c
r = 125 m
c
r = 150 m
c
r = 250 m
c
r = 500 m
c
m
μ
/
)
r
(
P
00
006
00
004
00
002
001
0
001
0
Experiment
Simulation
Analytics
20
40
Distance r from the center / µm
60
80
100
100
200
400
Distance r from the center / µm
300
500
m-
μ
/
)
r
(
P
0
FIG. 3. Radial probability densities P (r) for compartment
sizes, rc, between 25 µm and 500 µm. Each curve repre-
sents statistically averaged distributions from 2–10 indepen-
dent measurements. As rc increases, the maximum of P (r)
close to the compartment wall decreases monotonously. The
inset displays a close-up of the experimental data (7 indepen-
dent experiments, solid line), Brownian dynamics simulation
(dashed line) and analytical solution (dotted line) for rc =
100 µm. The shaded background displays the standard devi-
ation of the experimental data. See Materials and Methods
for details.
possible zone of flagella-wall contact interactions. The
maximum of P (r) decreases for increasing compartment
size from rc = 25 µm to rc = 500 µm, while the overall
shape of P (r) described above is preserved.
We also note that towards the compartment's center,
where there is no significant influence of the wall on the
swimming behavior, the alga swims with a typical veloc-
ity of 100±10 µm/s. This is in agreement with swimming
velocities reported in bulk [31]. As a consequence of
alga-wall interactions, the measured velocity slows down
to 60±20 µm/s within the near-wall swimming zone.
Brownian Dynamics Simulations
We compared these experimental results to Brownian
dynamics simulations adapted as follows. The Chlamy-
domonas cell is modeled as an active asymmetric dumb-
bell consisting of two rigid spheres, inspired by the model
developed in [32]. The smaller sphere represents the cell's
body and the larger sphere mimics the stroke-averaged
area covered by the beating of the two anterior flagella
(see Fig. 4B). An asymmetric dumbbell is the simplest
model for a cell to experience a torque during an inter-
action event with the confining wall. This torque is a
major ingredient in our simulations, since it may reori-
ent the alga away from the interface. All geometric and
dynamic parameters that entered the simulations were ei-
ther measured directly from our experiments or extracted
from the literature. This includes a finite microscopic in-
teraction time at the interface, which was inferred from
image sequences in [29]. In the simulations, we also con-
sider stochastic noise to account for the biological nature
of our microswimmer. However, the comparison to the
analytical approach (provided in the SM ) suggests that
this noise represents only a minor contribution. Hydro-
dynamic interactions are entirely absent in this model
and the dynamics are solely determined by the repulsive
force and torques at the confining wall (see Materials and
Methods).
The radial probability densities P (r) were extracted
from both simulations (see Fig. S1 for heat maps) and
analytical theory; exemplar curves are presented in the
inset of Fig. 3.
In general, we find good quantitative
agreement of these data with the experiment, namely an
increased likelihood of finding the alga located near the
interface for small compartments.
Near-Wall Swimming Probability
We define the near-wall swimming probability, Φ(rc),
as the relative probability of finding the alga/dumbbell
towards the wall as compared to the center. In our no-
4
500
400
300
200
100
0
0
(D)
r
m
c =
r
c =
m
r
m
c =
r
c
m
F
1200
1000
800
600
400
200
(E)
r
c
m
C
W
2
4
6
8
10
Detention time τ
/ s
0
0
20
40
80
60
θ
FIG. 4. Statistics of near-wall swimming.
(A) Near-wall swimming probability Φ(rc) for compartment radii ranging from
25 µm to 500 µm: experimental data (circles) denote mean values averaged over multiple independent experiments (crosses),
Brownian dynamics simulations (diamonds) and analytical calculation (solid line).
(Inset) A log-log representation of the
same data suggests a 1/rc scaling for sufficiently large compartments, i.e. rc ≥ 100 µm. (B) Asymmetric dumbbell model
representing the Chlamydomonas alga, see Materials and Methods for details. (C) Schematic illustration of the swimming angle
with respect to the wall tangential. (D) Detention time distributions for swimming near the concave interface for rc = 50 µm
and rc = 500 µm. (E) Experimental swimming angle distributions for rc = 50 µm near and far ("center") from the wall of the
compartment. The inset illustrates the definition of "wall" and "center" for the purposes of the distributions.
tation, this is written as:
Φ(rc) = 1 −
rc
rc − b Z rc
0
−b
P (r)dr
(2)
where b is the extent of wall influence, measured from
experiments as approximately 15 µm and independent of
compartment size. This distance b defines the near-wall
swimming region used later in our study.
Figure 4A presents Φ(rc) for experiments, Brownian
dynamics simulations, and analytical theory, which all
agree quantitatively and show a monotonic decrease for
increasing compartment radius rc. Furthermore, the data
suggest a linear scaling of Φ(rc) with the wall curvature,
i.e. 1/rc (see Fig. 4A inset). The deviation from this
linear scaling for strong confinement (rc < 100 µm) is
consistent with the fact that the alga experiences a torque
upon wall interaction that may allow it to occasionally
escape, even for highly curved walls.
Temporal and Angular Statistics of Near-wall
Swimming
We define the wall detention time τ as the time that
the alga spends within the near-wall swimming zone (dis-
tance b from the wall). As depicted in Fig. 4D, the
distribution of experimental detention times exhibits a
maximum at approximately τmax = 0.5 s for rc = 50 µm,
featuring an exponential decay e−t/τ ∗
towards longer de-
tention times with a decay time of τ ∗ = 1 s. For larger
compartments, the maxima of the detention time distri-
butions are shifted to shorter times and the decay times
are also significantly reduced (see Fig. 4D): τmax < 0.2 s
and τ ∗ = 0.3 s are found for rc = 500 µm, indicating
that the wall detention time becomes comparable to the
typical interaction time reported for flat interfaces [29].
Since the cell has a non-zero velocity at all times during
experiments, these results imply that the alga tends to
spend more time swimming near the interface in smaller
compartments than for larger ones. However, the alga's
detention time at the interface does not yield any infor-
mation about the directionality of the alga's swimming.
It also does not allow us to distinguish between two possi-
ble extreme cases: (i) the alga probes the wall repeatedly
at the same location, eventually escaping the interface
after some time, and (ii) the alga swims non-stop either
clock- or counterclockwise along the curved interface.
In order to characterize the swimming direction, we an-
alyzed the local swimming angle θ (see Fig. 4C), which is
measured relative to the local wall tangent. We consider
two regions in the compartment: "wall" and "center"
(see Fig. 4E inset). As discussed above, the near-wall re-
gion is defined from a distance b from the interface for all
compartment sizes. All trajectories outside this region
contribute to the swimming angle distribution towards
the center of the compartment. As shown in Figure 4E,
we find that θ displays an isotropic distribution towards
the center of the compartment. However, within the
near-wall swimming zone, θ shows a maximum around
zero swimming angle, indicating that the wall induces a
preferred swimming direction parallel to the concave in-
terface. The distribution decays towards larger angles
in line with Gaussian statistics (standard deviation 12◦,
solid line in Fig. 4E). Note that these experimental distri-
butions represent all navigational movement of the alga,
which may include any microscopic interactions the alga
might have with the compartment boundaries.
Based on the aforementioned analysis, the alga un-
dergoes the following process in isolated microcompart-
ments: Upon interaction with an interface, the alga re-
orients due to its characteristic torque and scatters off at
some shallow angle (see also [29]). It subsequently con-
tinues swimming predominantly ballistically in the com-
partment. If the compartment is sufficiently curved, the
alga will encounter another section of the interface in
a short time, interact, scatter off, and continue swim-
ming. This process will repeat itself such that for small
compartments (high curvature), it appears that the alga
swims non-stop parallel to the interface (see movie S1),
since the alga will encounter another interface during its
characteristic persistent swimming time. In contrast, for
large compartments (low curvature), the alga will travel
farther before meeting another interface. Thus, it is more
likely that the alga's reorientation will direct it towards
the compartment center. Nonetheless, due to the con-
finement the alga will encounter an interface before un-
dergoing any run-and-tumble-like motion. Our Brownian
dynamics simulations and analytical theory capture this
process: using an asymmetric dumbbell model, the alga
will naturally experience a torque at the interface and
reorient with a finite interaction time, subsequently en-
countering another interface before it can "tumble".
Geometries with Convex Interfaces
The results described above suggest that the alga may
not remain swimming near the convex interface of a cylin-
drical obstacle, in contrast to pusher-type swimmers, e.g.
bacteria [33] and spermatozoa [34]. An alga contained in
a circular compartment including a central pillar estab-
lishes a simultaneous comparison between the behavior
at concave and convex interfaces within the same com-
partment. The analysis of experimental and simulated
trajectories (see Fig. S2) show that the alga scatters off
at the pillar and escapes the convex wall. This observa-
tion is consistent with studies on single microscopic scat-
tering events at flat [29] and convex [30] interfaces. We
also note that P (r) in the vicinity of the convex pillar
interface is slightly increased as compared to the value
of P (r) away from any walls (see Fig. S2), which can
be attributed to the alga's finite interaction time at the
interface.
Motility in Elliptical Compartments
Finally, we consider elliptical chambers, which are
compartments possessing a curvature gradient along its
(A)
(B)
(C)
Simulations
!"#$%&'
8
6
4
2
5
Expt.
Sims.
Experiment
()*+,!+&'
0
0.01
0.02
0.03
0.04
κ / µm
FIG. 5. Motility in elliptical compartments. Relative proba-
bility density heat maps extracted from experiments (A) and
Brownian dynamics simulations (B) both display an enhanced
probability density in the apex regions of the elliptical com-
partment (eccentricity = 0.91). (C) The local maximum rela-
tive probability density in both experiments and simulations
scales with the local wall curvature κ. Lines represent linear
fits to the experimental (straight) and simulation (dashed)
data for wall curvatures κ ≤ 0.01µm−1.
wall. Experiments (Fig. 5A) and Brownian dynamics
simulations (Fig. 5B) show a higher likelihood of find-
ing the alga in one of the apex regions of the compart-
ments. We find in both experiments and simulations
that the local maximum of the relative probability den-
sity in elliptical compartments scales linearly with the
corresponding local wall curvature κ (Fig. 5C), in line
with the results obtained for circular compartments (see
Fig. 4A). We observe a deviation from this linear scaling
in both experiments and simulations for higher wall cur-
vatures, as also seen for circular compartments. Hence,
we have established unambiguous evidence that the (lo-
cal) wall curvature controls the near-wall swimming ef-
fect in confinement and that the results described for
circular compartments cannot be attributed to a trivial
surface-to-volume effect (see [35]). Note that a torque-
free, spherical active Brownian particle cannot reproduce
the experimental data (see Fig. S4). Thus, we find that
Chlamydomonas' motility in confinement is governed by
reorientations of the cell due to steric wall interactions
and the cell's characteristic torque.
DISCUSSION
In the absence of external flow, cell-cell interactions,
photo- and chemotaxes, we isolated a curvature-guided
motility mechanism for a single microalga in a confined
microfluidic habitat with controlled geometric properties.
The concave nature of the confining walls leads to an en-
hanced probability of near-wall swimming for puller-type
microswimmers, as quantified by a statistical analysis of
experimental cell trajectories. Brownian dynamics simu-
lations based on an active asymmetric dumbbell model,
as well as analytical theory, quantitatively capture the
experiments and validate a characteristic curvature scal-
ing of the near-wall swimming probability. The main in-
gredients of this curvature guidance are the torque that
the alga experiences during an interaction event with the
wall, the compartment's wall curvature, and the suppres-
sion of the alga's diffusive swimming regime in confine-
ment. We highlight that hydrodynamics are not explic-
itly necessary to understand this swimming behavior, yet
they might be required to capture the microscopic de-
tails of flagellar interactions with interfaces, as previously
reported [29, 30]. Our results suggest that enhanced
near-wall swimming in confinement is not exclusive to
microorganisms propelling themselves by rear-mounted
appendages, but is also manifested by microorganisms
with anterior flagella, such a microalga.
These results pave the way towards a fundamental un-
derstanding of the motility of microorganisms in their
natural habitats, which typically constitute a plethora
of curved interfaces. Chlamydomonas is a soil-dwelling
microorganism and represents one of the prime model or-
ganisms to study cellular processes, e.g. ciliary functions
in eukaryotes and photosynthesis, with a wide range of
applications including the synthesis of therapeutic pro-
teins and biofuels. In light of the idea that surface associ-
ation represents an efficient way of harboring a favorable
environment rather than being swept away by a liquid
current [16], the observed swimming behavior might rep-
resent an advantage in the cell's natural habitat.
From a technological perspective, we anticipate that
these insights may inspire new design principles for the
guidance of cellular motion, e.g. for the targeted delivery
of (pharmacological) cargo [36] and for directing motile
eukaryotic cells through microfluidic channels in bioengi-
neering applications [37–39], complementary to existing
rectification approaches [29]. Moreover, an immediate
consequence of enhanced detention times at a highly
curved wall is a greater likelihood for the adhesion of
planktonic cells at walls, which can trigger the coloniza-
tion and formation of biofilms in liquid-immersed porous
6
media. In such heterogeneous environments, biofilm for-
mation has substantial implications on transport pro-
cesses and permeability [40]. Thus, we expect that these
insights are highly relevant in environmental applications
related to bioremediation of contaminated groundwater
and soil, water filtration systems, and bioreactors for the
production of biofuels based on photosynthetic microal-
gae [41].
MATERIALS AND METHODS
Cell Cultivation
Cultures of wild-type Chlamydomonas
reinhardtii
(SAG 11-32b) were cultivated axenically in Tris-Acetate-
Phosphate (TAP) medium on a 12 h–12 h day-night cycle,
with daytime temperature of 24 ◦C and nighttime tem-
perature of 22 ◦C in a Memmert IPP 100Plus incubator.
The daytime light intensity was held at 1000–2000 Lux,
and reduced to 0% during the night. All experiments
were performed at the same time in the cell's life cycle in
order to ensure consistency in the cell's size and behavior.
In preparation for experiments, 15 mL of cell suspension
was centrifuged for 10 minutes at 100 g at room tempera-
ture; 10-13 mL of solution was subsequently removed and
the remaining 2-5 mL suspension was allowed to relax for
60-90 minutes. This suspension was diluted with room
temperature TAP to enhance the likelihood for capturing
precisely one cell in an isolated microcompartment.
Microfluidics
Arrays of stand-alone (i.e. no inlet or outlet) circu-
lar microfluidic compartments with a height of 20 µm
were created using standard PDMS-based soft lithogra-
phy techniques in a cleanroom. Additional experiments
were conducted with circular compartments containing a
pillar located at the center of the compartment, as well
as elliptical chambers. Prior to experiments, both the
PDMS device and a glass microscope slide were cleaned
using air plasma (Electronic Diener Pico plasma system,
100% exposure, 30 seconds). After plasma cleaning, a
small amount of 8wt% polyethelyne glycol was gently
rinsed over both surfaces to prevent adhesion of the alga
to the surfaces. After placing a droplet of the diluted
algal suspension onto the feature side of the PDMS, the
glass slide was placed on top and gently pressed to seal
the compartment. Only compartments containing pre-
cisely one cell were used for experiments.
Microscopy
Cell imaging was conducted using an Olympus IX-81
inverted microscope contained in a closed box on a pas-
sive anti-vibration table with an interference bandpass
filter (λ ≥ 671 nm, full-width-half-maximum of 10 nm)
in order to avoid any photoactive response of the cell.
Videos ranging from 5-30 minutes were recorded using a
Canon 600D camera at a frame rate of 24 frames per
second at full resolution (1920 pxl×1080 pxl). This corre-
sponds to approximately 7×103–40×103 total trajectory
points for a single experiment. The single-cell experi-
ments were repeated 2–10 times for each compartment
size, corresponding to 35×103–120×103 total trajectory
points.
Image Processing and Particle Tracking
The videos were sequenced into 8-bit grayscale images
with improved contrast using custom-made Matlab al-
gorithms. The compartment boundaries were manually
defined in order to denote the region of interest, as well
as the compartment's center, for particle tracking. Two-
dimensional particle detection was performed using al-
gorithms written for colloidal systems; particle tracking
was subsequently done in Matlab based on tracking al-
gorithms developed by Crocker and Grier [42].
Data Analysis
Mean-squared displacements (MSD) were extracted
using a custom-made Matlab script, based on trajecto-
ries containing a minimum of 7×103–14×103 data points
for the rc = 25 − 250 µm and at least 40×103 data points
for the rc = 500 µm. This corresponds to MSD curves
containing 2×103–10×103 points each.
A custom-made Matlab algorithm based on a pixel
grouping method for the data binning was applied to the
trajectory data to compute the relative probability den-
sity heat maps by c(x, y) = nbin/(AfracΣnbin), where nbin
is the number of trajectory points within bin area Abin,
and Afrac = Abin/Achamber is a geometric normalization
factor. The radial probability densities P (r) of the alga's
distance r from the center of the compartment were cal-
0 P (r)dr = 1. We define
culated using Eqn. 1 such that R rc
the near-wall swimming probability as in Eqn. 2.
Brownian Dynamics Simulations: Details of Steric
Wall Interactions
The force acting on the dumbbell at the wall is given
by ~Fw = ~F1 + ~F2 with ~Fα = −~∇Uα(r), α = 1, 2. We
use the Weeks–Chandler–Anderson repulsive potential
7
−(cid:0) aα
d (cid:1)6i + ǫ, if d < 21/6aα,
Uα(d)/(kBT ) = 4ǫh(cid:0) aα
d (cid:1)12
and 0 otherwise, where d is the distance of the sphere
α ∈ {1, 2} to the wall of the compartment, a1 = 5 µm,
a2 = 2.5 µm are the radii of the spheres (see Fig. 4B)
and ǫ = 10 is chosen to achieve a strong screening. Fur-
thermore, the torque is given by ~Tw = ~T1 + ~T2 where
~T1 = (~r1 − ~r) × ~F1 = l(~e × ~F1)/2, ~T2 = −l(~e × ~F2)/2,
and l = 5 µm. The position ~r of the dumbbell's cen-
ter of mass has the following equation of motion d~r
dt =
v0~e + γw ~Fw + ~η. Here, ~Fw is the steric wall interac-
tion and ~η is a Gaussian white noise with zero mean and
h~η(t)~η(t′)i = 2kBT γw1δ(t − t′). We use v0 = 60 µm/s
and kBT γw = 20 µm2/s, both based on experimental
measurements. Furthermore, the cell swims in the di-
rection ~e represented by a versor (a unit vector) pointing
from the second to the first sphere (see Fig. 4B). The
dt = ( ~Tw/τw + ~ξ) × ~e,
orientational equation of motion is d~e
where ~Tw is the torque acting at the wall and ~ξ is a
Gaussian white noise with h~ξ(t)~ξ(t′)i = 2kBT
1δ(t − t′),
τp
representing the tumble motion of the cell. Note that the
shear time at the wall τw = 0.15 s (extracted from [29])
and the persistence time τp = 5.1 s (taken from [31]) of
synchronous flagella beating are not connected via the
fluctuation-dissipation theorem. This is motivated by the
fact that the tumble time is associated with the active
motion of the cell, whereas the shear time is connected
to the interactions between wall and cell. Our simulated
system quantitatively reproduces the maximum of the
microscopic scattering angle distribution, found experi-
mentally in [29] (see Fig. S3).
∗ To whom correspondence should be addressed. E-mail:
[email protected]
[1] X. Sun, M.K. Driscoll, C. Guven, S. Das, C.A. Parent,
J.T. Fourkas, W. Losert, Asymmetric nanotopography
biases cytoskeletal dynamics and promotes unidirectional
cell guidance. Proc. Natl Acad. Sci. USA 112, 12557–
12562 (2015).
[2] H. Wu, M. Thi´ebaud, W.-F. Hu, A. Farutin, S. Rafaı,
M.-C. Lai, P. Peyla, C. Misbah, Amoeboid motion in
confined geometry. Phys. Rev. E 92, 050701 (2015).
[3] H. Jeon, S. Koo, W.M. Reese, P. Loskill, C.P. Grig-
oropoulos, K.E. Healy, Directing cell migration and or-
ganization via nanocrater-patterned cell-repellent inter-
faces. Nature Materials 14, 918–923 (2015).
[4] H. G. Yevick, G. Duclos, I. Bonnet, P. Silberzan, Archi-
tecture and migration of an epithelium on a cylindrical
wire. Proc. Natl Acad. Sci. USA 112, 5944–5949 (2015).
[5] M. Delarue, J. Hartung, C. Schreck, P. Gniewek, L. Hu,
S. Herminghaus, O. Hallatschek, Self-driven jamming in
growing microbial populations. Nature Physics 12, 762–
766 (2016).
[6] H. C. Berg, L. Turner, Chemotaxis of bacteria in glass
capillary arrays. Escherichia coli, motility, microchan-
nel plate, and light scattering. Biophys. J. 58, 919–930
(1990).
[7] R. U. Meckenstock, F. von Netzer, C. Stumpp, T. Lued-
ers, A.M. Himmelberg, N. Hertkorn, P. Schmitt-Kopplin,
M. Harir, R. Hosein, S. Haque, D. Schulze-Makuch, Wa-
ter droplets in oil are microhabitats for microbial life.
Science 345, 673–676 (2014).
[8] J. Wierzchos, A. de los R´ıos, C. Ascaso, Microorganisms
in desert rocks: the edge of life on Earth. Int. Microbiol.
15, 171–181 (2012).
[9] C. K. Robinson, J. Wierzchos, C. Black, A. Crits-
Christoph, B. Ma, J. Ravel, C. Ascaso, O. Artieda, S.
Valea, M. Rold´an, B. G´omez-Silva, J. DiRuggiero, Mi-
crobial diversity and the presence of algae in halite en-
dolithic communities are correlated to atmospheric mois-
ture in the hyper-arid zone of the Atacama Desert. Env.
Microbiol. 17, 299–315 (2015).
[10] L. Ranjard, A. Richaume, Quantitative and qualitative
microscale distribution of bacteria in soil. Research in
Microbiology 152, 707 – 716 (2001).
[11] M. Eisenbach, L. C. Giojalas,
Sperm guidance in
mammals-an unpaved road to the egg. Nature Reviews
Molecular Cell Biology 7, 276–285 (2006).
[12] P. Denissenko, V. Kantsler, D. J. Smith, J. Kirkman-
Brown, Human spermatozoa migration in microchannels
reveals boundary-following navigation. Proc. Natl Acad.
Sci. USA 109, 8007 (2012).
[13] V. Kantsler, J. Dunkel, M. Blayney, R. E. Goldstein,
Rheotaxis facilitates upstream navigation of mammalian
sperm cells. eLIFE 3, e02403 (2014).
[14] R. Nosrati, A. Driouchi, C. M. Yip, D. Sinton, Two-
dimensional slither swimming of sperm within a mi-
crometre of a surface. Nature Communications 6, 8703
(2015).
[15] N. Heddergott, T. Kruger, S.B. Babu, A. Wei, E. Stel-
lamanns, S. Uppaluri, T. Pfohl, H. Stark, M. Engstler,
Trypanosome motion represents an adaptation to the
crowded environment of the vertebrate bloodstream.
PLoS Pathog 8, 1–17 (2012).
[16] P. Watnick, R. Kolter, Biofilm, City of Microbes. Journal
of Bacteriology 182, 2675–2679 (2000).
[17] L. Hall-Stoodley, J. W. Costerton, P. Stoodley, Bacte-
rial biofilms: from the natural environment to infectious
diseases. Nat. Rev. Microbiol. 2, 95–108 (2004).
[18] H.-C. Flemming, J. Wingender, The biofilm matrix. Nat.
Rev. Microbiol. 8, 623–633 (2010).
[19] M. G. Mazza, The physics of biofilms–an introduction.
J. Phys. D: Appl. Phys. 49, 203001 (2016).
[20] E. Lauga, T. R. Powers, The hydrodynamics of swim-
ming microorganisms. Rep. Prog. Phys. 72, 096601
(2009).
[21] T. Brotto, J.-B. Caussin, E. Lauga, D. Bartolo, Hydro-
dynamics of confined active fluids. Phys. Rev. Lett. 110,
038101 (2013).
[22] J. Elgeti, R. G. Winkler, G. Gompper, Physics of
microswimmers-single particle motion and collective
behavior: a review. Rep. Prog. Phys. 78, 056601 (2015).
[23] K. Drescher, J. Dunkel, L. H. Cisneros, S. Ganguly, R. E.
Goldstein, Fluid dynamics and noise in bacterial cell–cell
and cell–surface scattering. Proc. Natl Acad. Sci. USA
108, 10940–10945 (2011).
[24] K. Drescher, R. E. Goldstein, N. Michel, M. Polin, I.
Tuval, Direct measurement of the flow field around
swimming microorganisms. Phys. Rev. Lett. 105, 168101
(2010).
8
[25] J. S. Guasto, K. A. Johnson, J. P. Gollub, Oscillatory
flows induced by microorganisms swimming in two di-
mensions. Phys. Rev. Lett. 105, 168102 (2010).
[26] K. C. Leptos, J. S. Guasto, J. P. Gollub, A. I. Pesci,
R. E. Goldstein, Dynamics of enhanced tracer diffusion
in suspensions of swimming eukaryotic microorganisms.
Phys. Rev. Lett. 103, 198103 (2009).
[27] Rothschild. Non-random distribution of bull spermato-
zoa in a drop of sperm suspension. Nature 198, 1221
(1963).
[28] P. D. Frymier, R. M. Ford, H. C. Berg, P. T. Cummings,
Three-dimensional tracking of motile bacteria near a solid
planar surface. Proc. Natl Acad. Sci. USA 92, 6195–6199
(1995).
[29] V. Kantsler, J. Dunkel, M. Polin, R. E. Goldstein, Cil-
iary contact interactions dominate surface scattering of
swimming eukaryotes. Proc. Natl Acad. Sci. USA 110,
1187–1192 (2013).
[30] M. Contino, E. Lushi, I. Tuval, V. Kantsler, M. Polin, Mi-
croalgae scatter off solid surfaces by hydrodynamic and
contact forces. Phys. Rev. Lett. 115, 258102 (2015).
[31] M. Polin, I. Tuval, K. Drescher, J. P. Gollub, R. E. Gold-
stein, Chlamydomonas swims with two "gears" in a eu-
karyotic version of run-and-tumble locomotion. Science
325, 487–490 (2009).
[32] A. Wysocki, J. Elgeti, G. Gompper, Giant adsorption of
microswimmers: Duality of shape asymmetry and wall
curvature. Phys. Rev. E 91, 050302(R) (2015).
[33] S. E. Spagnolie, G. R. Moreno-Flores, D. Bartolo, E.
Lauga, Geometric capture and escape of a microswimmer
colliding with an obstacle. Soft Matter 11, 3396–3411
(2015).
[34] D. Takagi, J. Palacci, A. B. Braunschweig, M. J. Shelley,
J. Zhang, Hydrodynamic capture of microswimmers into
sphere-bound orbits. Soft Matter 10, 1784–1789 (2014).
[35] Y. Fily, A. Baskaran, M.F. Hagan, Dynamics of self-
propelled particles under strong confinement. Soft Matter
8, 3002–3009 (2014).
[36] D. B. Weibel, P. Garstecki, D. Ryan, W.R. DiLuzio, M.
Mayer, J.E. Seto, G.M. Whitesides, Microoxen: microor-
ganisms to move microscale loads. Proc. Natl Acad. Sci.
USA 102, 11963–11967 (2005).
[37] X. Ai, Q. Liang, M. Luo, K. Zhang, J. Pan, G. Luo,
Controlling gas/liquid exchange using microfluidics for
real-time monitoring of flagellar length in living Chlamy-
domonas at the single-cell level. Lab on a Chip 12, 4516–
4522 (2012).
[38] S. K. Min, G. H. Yoon, J. H. Joo, S. J. Sim, H. S.
Shin, Mechanosensitive physiology of Chlamydomonas
reinhardtii under direct membrane distortion. Scientific
Reports 4, 4675 (2014).
[39] S. Das, A. Garg, A.I. Campbell, J. Howse, A. Sen, D.
Velegol, R. Golestanian, S.J. Ebbens, Boundaries can
steer active Janus spheres. Nature Communications 6,
8999 (2015).
[40] A. B. Cunningham, W. G. Characklis, F. Abedeen, D.
Crawford,
Influence of biofilm accumulation on porous
media hydrodynamics. Environmental Science & Tech-
nology 25, 1305–1311 (1991).
[41] P. M. Schenk, S.R. Thomas-Hall, E. Stephens, U.C.
Marx, J.H. Mussgnug, C. Posten, O. Kruse, B. Han-
kamer, Second generation biofuels: high-efficiency mi-
croalgae for biodiesel production. Bioenerg. Res. 1, 20–43
(2008).
[42] J. C. Crocker, D. G. Grier, Methods of digital video
microscopy for colloidal studies. J. Colloid Interface Sci.
179, 298–310 (1996).
ACKNOWLEDGEMENTS
General: The authors gratefully acknowledge M.
Lorenz and the Algae Culture Collection (SAG) in
Gottingen, Germany, for providing the Chlamydomonas
reinhardtii strain SAG 11-32b. We also thank S. Her-
minghaus for insightful discussions, as well as A. Schella
and D. Lavrentovich for discussions and technical assis-
9
tance. Funding: F.S and M.G.M. acknowledge finan-
cial support from the DFG Collaborative Research Cen-
ter SFB 937 (Project A20). O.B. acknowledges support
from the ESPCI Joliot Chair. Author contributions:
M.G.M. and O.B. designed research; T.O., F.S., T.B.,
C.K., J.C., M.G.M. and O.B. performed research; T.O.,
F.S., T.B., C.K., J.C., M.G.M. and O.B. analyzed data;
T.O., F.S., M.G.M and O.B. wrote the paper. Compet-
ing interests: The authors declare that they have no
competing interests. Data and materials availabil-
ity: All data needed to evaluate the conclusions in the
paper are present in the paper and/or the Supplementary
Materials. Additional data related to this paper may be
requested from the authors.
|
1311.4812 | 1 | 1311 | 2013-11-19T17:43:43 | Lattice solution model for order-disorder transitions in membranes and Langmuir monolayers | [
"physics.bio-ph",
"cond-mat.soft",
"cond-mat.stat-mech",
"q-bio.BM"
] | Lipid monolayers and bilayers have been used as experimental models for the investigation of membrane thermal transitions. The main transition takes place near ambient temperatures for several lipids and reflects the order-disorder transition of lipid hydrocarbonic chains, which is accompanied by a small density gap. Equivalence between the transitions in the two systems has been argued by several authors. The two-state statistical model adopted by numerous authors for different properties of the membrane, such as permeability, diffusion, mixture or insertion of cholesterol or protein, is inadequate for the description of charged membranes, since it lacks a proper description of surface density. We propose a lattice solution model which adds interactions with water molecules to lipid-lipid interactions and obtain its thermal properties under a mean-field approach. Density variations, although concomitant with chain order variations, are independent of the latter. The model presents both chain order and gas-liquid transitions, and extends the range of applicability of previous models, yielding Langmuir isotherms in the full range of pressures and areas. | physics.bio-ph | physics | Lattice solution model for order-disorder transitions in
membranes and Langmuir monolayers
Henrique S. Guidi∗ and Vera B. Henriques†
Instituto de F´ısica da Universidade de Sao Paulo
(Dated: February 28, 2018)
Abstract
Lipid monolayers and bilayers have been used as experimental models for the investigation of
membrane thermal transitions. The main transition takes place near ambient temperatures for
several lipids and reflects the order-disorder transition of lipid hydrocarbonic chains, which is
accompanied by a small density gap. Equivalence between the transitions in the two systems has
been argued by several authors. The two-state statistical model adopted by numerous authors
for different properties of the membrane, such as permeability, diffusion, mixture or insertion of
cholesterol or protein, is inadequate for the description of charged membranes, since it lacks a
proper description of surface density. We propose a lattice solution model which adds interactions
with water molecules to lipid-lipid interactions and obtain its thermal properties under a mean-field
approach. Density variations, although concomitant with chain order variations, are independent
of the latter. The model presents both chain order and gas-liquid transitions, and extends the range
of applicability of previous models, yielding Langmuir isotherms in the full range of pressures and
areas.
3
1
0
2
v
o
N
9
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
2
1
8
4
.
1
1
3
1
:
v
i
X
r
a
∗ E-mail me at: [email protected]
† E-mail me at: [email protected]
1
I.
INTRODUCTION
Lipid monolayers [1, 2] and bilayers [3, 4] have been extensively used as experimental
models for the investigation of thermal and structural properties of the biological membrane.
Phospholipid molecules form an ordered monolayer film on the air-water interface, with lipid
headgroups resting on water, while lipid hydrophobic chains acquire approximately parallel
orientation in the air phase, thus avoiding contact with water. External lateral pressure
guarantees aggregation into the monolayer film.
In water solution, lipids aggregate into
bilayer vesicles, as polar heads shield hydrocarbonic tails from contact with the aqueous
medium. Bilayers are tension-free and aggregation is driven by the hydrophobic effect.
Both systems may undergo several phase transitions. One of the most thoroughly inves-
tigated is the pronounced order-disorder lipid chain transition, with latent heat, presented
by either system. For lipid membranes temperature or pH variations may yield a so called
main gel-fluid transition. In the case of lipid monolayers, compression or heating disclose a
transition traditionally known as a liquid-condensed liquid-expanded transition. An abrupt
variation in lipid surface density is accompanied by disordering of the hydrocarbon chains.
The latter acquire "kinks", while distance between polar headgroups increases, yielding de-
creased surface density. This is recognized as the chief effect also for bilayers [5 -- 7], which
undergo the main transition under temperature-variation.
There are different advantages in adopting one experimental model or the other. Direct
measurement of surface area per lipid molecule, whose abrupt variation signals the transi-
tion, is possible only for monolayers. However, the discontinuity in lipid area depends on
external applied pressure. Equivalence between the two systems for a particular pressure on
monolayers has been argued by many authors [8 -- 10] and arguments rest on the assumption
of negligible interaction between the two leaflets that compose the bilayer. However, recent
studies indicate that this might be too strong a hypothesis [4]. Thus differences should be
subject of further investigation.
Lipids on the air-water interface reduce the surface tension, as lipid headgroups disrupt
the specially stable hydrogen bonds between surface water molecules [11]. Bilayers are free of
internal surface pressure. The hydrophobic effect, which consists of 'micro' phase-separation
between lipids and water, as a consequence of the disparity between the values of Van der
Waals attraction between aliphatic chains as compared to the value of water hydrogen bond
2
energies, is considered to be the main force driving membrane behaviour. This effect is
absent in the case of monolayers, since chains avoid water by turning to the air subphase.
In this study, we approach some of the questions related to the main transition from the
point of view of a minimal statistical model. Different two-state and multi-state models,
inspired on the success of Ising-like model for magnetism, were proposed by several authors
in the 1970's-1980's [12 -- 14], reflecting the possible states of the lipid chains, either extended
(all-trans) or disordered (gauche kinks) by different degrees, with focus on the different
areas occupied by the lipid head-groups on the bilayer surface. The orientational order of
lipids in the layer also suggested inspiration on models for nematic liquid crystals. On the
other hand, the large enthalpy attributed to chain melting suggested that this would be the
main entropic mechanism for the transition, thus demanding accurate treatment of chain
configurations, an approach followed by Nagle [5].
Marcelja [7] proposed that chain kinks could be treated in terms of a nematic-like order
parameter along the chain, subject to an effective field due to density of extended chains.
Thus chain entropy was obtained from the statistical calculation. Area per lipid headgroup
was taken as linearly dependent on the inverse of lipid chain length, from molecular volume
conservation. Caille [12] considered a lattice of two-state particles, corresponding to the
ordered and disordered chains. The disordered chain lipid would occupy a certain number
of sites, and was attributed an intramolecular chain entropy. Some authors looked also at
multi-state models [12, 14]. A lattice-gas three-state model, with two states for the lipids,
was also suggested [15], but the authors left out the essential intra-chain degeneracy. Nagle
[5] proposed a different approach: chain configurations of a two-dimensional section of the
lipid monolayer could be exactly enumerated through mapping on a dimer counting problem.
However, the much simpler treatment of chain entropy in terms of an average degeneracy of
the two-states model came to dominate the literature.
Doniach [13] simplified the two-state model of [12] by associating to each a different area
parameter in an ad hoc fashion, and noticed the possibility of treating the resulting model
exactly, through mapping on the seminal two-dimensional Ising model. Doniach's version of
the original two-state model turned into the most successful statistical model for the main
transition of the lipid system.
Inspite of its success for the description of the transition
for multicomponent lipid membranes, for diffusive properties, or for the effect of protein
or cholesterol insertion [3, 14, 16 -- 18], the model focuses on the order-disorder transition of
3
the lipid chains, while the area per chain is introduced in an ad hoc manner, by attributing
different areas to the 'ordered' and 'disordered' sites. As a consequence, a true discontinuity
in density is not displayed by the model at the order-disorder transition of the chains. This
aspect represents an important limitation in the case of competing interactions, such as for
ionic lipid layers [19 -- 21]. For dissociating lipids, electrostatic repulsion competes with the
hydrophobic effect, and a delicate equilibrium is established between chain order, charge
and molecular surface density. In this case, precise description of the local lipid density is
essential in order to appropriately rationalize thermal, electrical and structural properties
of the experimental system [19, 22].
In this study we have considered a two-state lipid lattice solution in which sites may
be occupied either by lipid or by water particles. Chains may be in two different states,
one of them largely degenerate, but the density results from the equilibrium occupation of
the lattice. Thus area per lipid is obtained from the statistics of the model. Also, proper
treatment of pressure allows examination of the equivalence hypothesis of mono and bilayers.
While recognizing the fact that different authors contributed to the formulation of the
original two-state model, in the name of simplicity we shall identify our model as a lattice
solution Doniach model.
In section 2, we define the statistical model.
In section 3, we
present our mean-field approach. Results for thermodynamic properties and phase diagrams
are displayed in section 4 . Physical interpretation in terms of the two systems of interest,
monolayers and bilayers, is discussed in section 5. Final comments are in section 6.
II. DEFINITION OF THE STATISTICAL MODEL
We revise the seminal model proposed by Doniach a few decades ago [3, 13, 14] for the
phospholipid bilayer main transition in order to set notation. In Doniach's lattice model,
lipid chains fill the plane lattice and are considered to visit two different particle states, an
ordered chain state o (Fig. 1a), corresponding to an extended chain, and a highly degenerate
disordered chain state d, meant to represent an average shortened chain (Fig. 1b).
The system consists of N particles distributed over the L2 = No − Nd sites. Its configu-
rational energy may be written as
E = −ǫooNoo − ǫddNdd − ǫodNod ,
(1)
4
(a)
(b)
}
,
,
,
,
,
,
, ...,
Ω
{z
(c)
ǫoo
(d)
ǫod
(e)
ǫdd
FIG. 1: (a) Simplified representation of a lipid with extended hydrocarbon chains. (b)
Disordered chain configurations are represented through a single average disordered state
of degeneracy Ω. (c)-(e) interaction between pairs of lipids in different states.
where Nxy is the number of contacts between two particles in states x and y, and Nx is the
number of lipids in state x, where x = {o, d} and y = {o, d}. Interaction parameters ǫxy
should all be taken as positive, since they represent effective attraction between particles.
At the main transition, there is a sharp variation of the lipid chain states. A chain order
parameter
describes chain order.
m =
No − Nd
L2
(2)
The model incorporated a second feature. Disordering of the chains is intimately related
physically to loosening of the packing of lipid headgroups, as indicated by experiments. It
seemed natural to make lipid area dependent on lipid chain state. Thus in Doniach's proposal
a lipid particle in the ordered state could be associated with surface particle area ao, while
a lipid molecule in one of the disordered states would be given area ad, with ad > ao. As a
consequence, lattice and model areas have no correspondence, with A ≥ L2, while Nlip = L2.
This approach implies that the area per particle aDoniach is defined as
aDoniach =
Noao + Ndad
L2
.
(3)
As it can be seen, chain order parameter m and area per particle aDoniach are not independent
thermodynamic variables, since
aDoniach(m) =
1
2
m(ao − ad) +
1
2
(ao + ad) .
(4)
The model simplicity allows mapping on a modified form of the two-state Ising mag-
netic model, whose thermodynamic properties are very well established. Lateral pressure,
5
(a)
(b)
FIG. 2: (a) Illustration of Doniach's model. Lipid states define area per lipid, which is
independent of lattice spacings. (b) A lattice gas version of Doniach's model (DLG). Sites
are occupied either by lipid or water particles. Area per lipid is obtained from model
statistics.
conjugate to 'area', acts as en effective field favouring the ordered phases. Together with
temperature, this field controls the lipid system phases. Chains are ordered at large lateral
pressures and low temperatures, as expected. At fixed temperatures, chains order discontin-
uously under increasing lateral pressure. As temperature increases, the transition disappears
at a critical temperature.
Despite the utility of the model for many purposes, it may be unsuitable under certain
circumstances, such as in the case of charged lipid membranes [19], whose thermodynamic
properties depend strongly on charge surface density.
Doniach's lattice solution model - DLG
We propose to introduce lipid density as a true statistical variable and write Doniach's
model as a lattice solution with explicit water particles. The inclusion of interactions between
lipid and water particles is essential for the purpose of investigation of the bilayer-monolayer
analogy [8] and Marsh[10]. Like with its predecessor, our goal is to describe the order-
disorder transition of a surface of lipid chains within the framework of a simplified model.
In this new proposal, the fixed relation between area per particle and chain parameter
m, Eq. 2, is abandoned. Figure 2 illustrates pictorially our proposal, as compared to the
original Doniach description.
Let us consider a square lattice of area A and L2 = A sites, which may be occupied by
lipids or by water. We define occupational variables (δ = 1) and (δ = 0), for lipids and water,
6
TABLE I: Particle states and statistical variables
particle
occupation variable σ
ordered chain lipid (o)
disordered chain lipid (d)
water (w)
1
−1
0
mapping σ → η
1
2 σ2(1 + σ)
1
2 σ2(1 − σ)
1 − σ2
respectively. The lipid particle chains may be either in the ordered or disordered state, with
chain variables given by (η = 1) or (η = 0), accordingly. Interactions between lipid particles
are the same as those of Eq. 1, but the number of lipid particles Nlip = No + Nd is not fixed.
Model energy reads
E = −ǫooNoo − ǫodNod − ǫddNdd
−ǫowNow − ǫdwNdw − ǫwwNww ,
(5)
where Nxy is the number of contacts between sites occupied by lipids in ordered state o, by
lipids in disordered state d or by water particles w. As in Doniach's model, the disordered
states are multiply degenerate, the degeneracy Ω corresponding to the high entropy of the
disordered hydrocarbon chains of a single lipid.
In order to give our model a statistical treatment, it is convenient to rewrite energy (Eq.
5) in terms of statistical variables. We attribute variables σ to lattice sites as in table I.
Under this notation, Eq. 5 is rewritten as
E = Xx Xy
−ǫxyX(ij)
i ηy
ηx
j
(6)
where η is defined in table I.
Under this representation, and after some manipulation, system energy (Eq. 5) is given
by an interaction term Eint and a hydrophobic "field" term Ehydr, besides a constant term:
E = Eint + Ehydr + E0
= −J X(ij)
−KX(ij)
σiσj − ∆X(ij)
j + IXi
σ2
i σ2
σiσj (σi + σj)
σ2
i − 2ǫwwA ,
(7)
7
where, for simplicity, new interaction parameters are defined as
J =
ǫoo + ǫdd − 2ǫod
4
,
∆ =
ǫoo − ǫdd − 2ǫow + 2ǫdw
4
,
and
K =
ǫoo + ǫdd + 2ǫod + 4ǫww − 4ǫow − 4ǫdw
4
I = 2(2ǫww − ǫow − ǫdw) .
(8)
(9)
(10)
(11)
For the lattice solution lipid model, equilibrium properties are more easily calculated in
the grand-canonical ensemble. The grand-partition function reads
Ξ(T, µlip, µw) = X{σ}
ΩNd e−β(E−µlipNlip−µwNw) ,
(12)
where µlip and µw are the lipid and water chemical potentials. The total number of lipid
particles Nlip, the number of water particles Nw and the number of disordered chain lipids
Nd are given respectively by
and
and
Nlip = Xi
σ2
i
Nw = A − Nlip
Nd = Xi
σ2
i (1 − σi)/2 .
(13)
(14)
(15)
It is interesting, at this point, to note that the linear "hydrophobic" energy term in the
energy expression (Eq. 7) competes with the chemical potential factor, so we rewrite the
grand-partition function as
Ξ(T, µ) = eβ(µw+2ǫww)AX{σ}
ΩNd(σ) e−β(Eint({σ})−µNlip({σ}) ,
(16)
where µ ≡ µlip − µw − I.
8
III. MEAN-FIELD APPROACH
The model equilibrium properties may be obtained from a Curie-Weiss mean-field ap-
proach [23] which allows linearization of the system energy in the statistical sums and turns
exact calculation possible. Interactions are made independent of distance, with the replace-
ment
X(i,j)
XiXj →
q
2AXi
XiXj
Xj ,
(17)
where Xi are interaction variables, A = L2 is the system area and q is the model coordination
number. Under this transformation, the model energy is written as:
EMF =
q
2A (cid:18)−JM 2 − 2∆MN − KN 2 + IN
A
2 − ǫwwA2(cid:19) ,
where N is the global number of particles given by Eq. 13, and M is defined as
M = Xi
σi .
(18)
(19)
Linearization of the quadratic terms in M and N in the grand-partition function of Eq.
16, with model energy E, Eq. 7, replaced by mean-field energy, EMF (Eq. 18), may be
achieved through Gaussian transformations
ey2
=
1
√π Z ∞
−∞
e−x2+2yx dx .
(20)
Summation over statistical variables σ becomes straightforward. Integrals in the newly
introduced variables X are then solved by the steepest descent method. For large sys-
tems the main contribution comes from extrema which yield the following grand-potential
Φ(T, A, µ, H):
Φ(T, A, µ, H; m, n)
A
= 2(Jm2 + Kn2 + 2∆nm)
−2ǫww−µw −
1
β
ln (1 + φ+) .
(21)
Here m, n are respectively chain order parameter m = hMi /A and lipid densities n =
φ+
hNlipi /A. φ+ is a function of m and n, defined as
= eβ(4∆m+4Kn+µ− 1
eβ(4J m+4∆n)
+
−
φ−
2 I) ×
Ω e−β(4J m+4∆n)
.
9
(22)
The chain order parameter m(T, µ) and model density n(T, µ), are then given through
the following system of coupled equations which define the conditions for the extrema for
exponents of the Gaussian integrals:
m
n
(T, µ; m, n) =
φ−
φ+
1
1 + φ+
.
(23)
IV. MODEL THERMODYNAMIC PROPERTIES AND PHASE DIAGRAM
Model properties are investigated through inspection of the solutions for chain order
parameter m and lipid density n (Eqs. 22 and 23). Depending on the thermodynamic
parameters, several solutions may be found, and the equilibrium physical solution is obtained
from inspection of the global minimum of the grand-potential Φ. Fig. 3a-d illustrates this
procedure. Differently from the original Doniach model, it can be seen that the disordering
transition of the chains, signalled by the abrupt discontinuity in chain parameter m, is
accompanied by a small discontinuous transition in density n, shown in the detail.
A possible phase diagram is displayed in Fig. 4 for a specific set of parameters. Three
phases are present: a gas phase (Gas), characterized by very low density (n ≈ 0); a liquid of
ordered chains (Ord), with chain parameter m ≈ 1; and a liquid of disordered chains (Dis),
with chain parameter m ≈ −1. The gas phase is present at low chemical potentials. At low
fixed temperature, a Gas-Ord transition takes place as chemical potential is increased. For
high chemical potentials, as one increases temperature, a discontinuous Ord-Dis occurs, with
a small density gap in density n accompanying a sharp transition in chain order parameter
m, from 1 to −1. For a range of intermediate chemical potentials, raising temperature
produces a Gas-Dis discontinuous transition in density n. The three phases coexist at a
triple point.
Reentrant behaviour is displayed by the model system, near the triple point, in a small
range of chemical potentials. This is shown in the detail of Fig. 4: as temperature is
increased, the ordered chains give place to a gas phase, and as temperature is increased
further, the gas phase presents coexistence with a disordered chains liquid.
The different phases and phase transitions present in the phase diagram are illustrated
in Figs. 5 and 6.
In Fig. 5, chain parameter m and lipid density n behaviour with temperature, at fixed
10
(a)
0
0.2
0.4
0.6
t
0.8
1.0
1.2
1.0
0.5
m
0
−0.5
−1
1.0
0.8
0.6
0.4
0.2
0.0
n
φ
1
−1
−3
−5
−7
(b)
0.8
1.0
1.2
0.8
1.0
1.2
(d)
0
0.2
0.4
1.000
n
0.998
(c)
0
0.2
0.4
0.6
t
0.6
t
0
0.2
0.4
0.6
t
0.8
1.0
1.2
FIG. 3: Thermal behaviour of chain order parameter m (a) and of density n (b). A detail
of the discontinuity in density is displayed in (c), and behaviour of free-energy as a
function of temperature in (d). Dashed lines represent all the possible solutions for Eqs. 23
and 22. Continuous lines correspond to the absolute minimum of the free-energy.
chemical potential, are shown in different regions of the phase diagram. The Ord-Dis tran-
sition, with m going from +1 to −1, is again shown to be accompanied by a discontinuity
in density n, with increasing area per lipid. However, the discontinuity in density decreases
as chemical potential is raised (compare Figs. 5(d) and 5(e). The reentrant behaviour with
the sequence of transitions Ord-Gas and Gas-Dis as temperature is raised is illustrated in
11
c
7
b
7
a
7
0.8
t
0.6
Gas
Dis
8c
8b
8a
0.7
t
0.5
−6.45
µ
−6.30
Ord
0.4
−8
−7
−6
µ
−5
FIG. 4: Model phase diagram in the temperature t/J vs chemical potential µ/J plane.
Model parameters are ∆/J = 0.6, K/J = 1 and Ω = 1000. Continuous lines are
coexistence lines. Hollow circle is a triple point. The gas-disordered chains coexistence line
ends at a critical point indicated by a full circle. Dashed lines represent cuts illustrated in
Figs. 5 and 6.
Figs. 5(c) and 5(f): first chain parameter m goes discontinuously from 1 to 0, while density
jumps from 1 to 0, and at higher temperature the chain parameter goes from 0 to −1, while
density raises from 0 to near 80%. The different phase transitions of Fig. 5 are represented
as dashed lines in the phase diagram of Fig. 4.
Chain parameter m and lipid density n behaviour as chemical potential is raised, at
fixed temperature, is shown in Fig. 6, for different regions of the phase diagram. The Gas-
Ord transition at low temperatures is signaled by a discontinuity in m from 0 to 1, with
a density jump from ≈ 0.1 to 1 (Figs. 6(a) and 6(d)). At intermediate temperature, the
Gas-Dis transition is followed by a Dis-Ord transition, with two discontinuities in density
(Figs. 6(b) and 6(d)). Finally, at higher temperature, a Gas-Dis transition is accompanied
by a density jump between densities 0.2 and 0.8 (Figs. 6(c) and 6(e) ). The different phase
transitions of Fig. 6 are represented as dashed lines in the phase diagram of Fig. 4.
What is the effect of varying model parameters upon the phase diagram? Figure 7a
illustrates the effect of variation of parameter ∆ at fixed K, while Fig. 7b illustrates the effect
of varying K at fixed ∆. Inspection of role of the interaction parameters in the expression
for energy (Eq. 7) explain some of the features displayed by the different phase diagrams.
Parameter K favors site states σ = +1 and σ = −1 and thus the filling of the lattice by lipids,
at low temperatures. Thus, the lipid gas-liquid transition at low temperatures is moved
12
b
b
c
1.0
0.5
m
0
−0.5
−1
0
1.0
0.5
m
0
−0.5
−1
0
1.0
0.5
m
0
−0.5
−1
0
(a)
1.00
0.99
n
0.98
0.97
0.96
(d)
0.5
t
1.0
0
0.5
t
1.0
(b)
n
1.00
0.95
0.90
0.85
(e)
0.5
t
1.0
0
0.5
t
1.0
(c)
n
1.0
0.8
0.6
0.4
0.2
0
(f)
0.5
t
1.0
0
0.5
t
1.0
FIG. 5: Chain order parameter m (a-c) and lipid density n (d-e) as functions of
temperature t, for fixed chemical potentials µ −5, −6.2 and −6.39. Figs (a) and (d)
illustrate that at higher chemical potentials the discontinuous chain parameter transition is
accompanied by a small density discontinuity. Figs (b) and (d) display model behavior
nearer to the triple point, with a larger discontinuity of density at the order-disorder
transition. Figs (c) and (e) show reentrant behaviour beyond the triple point, with an
order-gas transition and a gas-disordered chains transition as temperature is raised. The
three cuts above are shown as dashed lines in the phase diagram of Fig. 4.
towards lower chemical potential µ, as K is increased. On the other hand, parameter ∆
favors particle state σ = +1, and thus stabilizes the the ordered chains liquid state, moving
the low temperature gas-liquid transition to lower chemical potential µ and the ordered
chain liquid - disordered chain liquid transition to higher temperature T . Dislocation of
the coexistence lines might yield the disappearance of the Gas-Dis line, and therefore of the
triple and critical points. This is the case both for K = 0.8 (∆ = 0.6) in Fig. 7a and for
∆ = 0.7 (K = 1) in Fig. 7.
13
1.0
0.5
m
0
−0.5
−1
1.0
0.5
m
0
−0.5
−1
1.0
0.5
m
0
−0.5
−1
(a)
1.0
0.8
0.6
n
0.4
0.2
0
(d)
−8
−7
µ
−6
−5
−8
−7
µ
−6
−5
(b)
1.0
0.8
0.6
n
0.4
0.2
0
(e)
−8
−7
µ
−6
−5
−8
−7
µ
−6
−5
(c)
1.0
0.8
0.6
n
0.4
0.2
0
(f)
−8
−7
µ
−6
−5
−8
−7
µ
−6
−5
FIG. 6: (a) Isothermal model transitions with chemical potential µ are illustrated in Figs
(a)-(e), for temperatures t 0.65, 0.684 and 0.71, in terms of chain order parameter m and
density n. (a) and (d) illustrate the gas-ordered chains transition below the triple point.
(b) and (e) show the double transition, gas-disordered and disordered-ordered chains,
respectively, just above the triple point. (c) and (f) display data for the gas-disordered
chains transition, further above the triple point. The three cuts are represented as dashed
lines in the phase diagram of Fig. 4.
Thus, the two coexistence lines, gas liquid and ordered-disordered liquid may either merge
continuously or meet at a triple point. But why does the coexistence line between the gas
and the disordered chain liquid disappear, as ∆ is increased at fixed K or as K is increased
at fixed ∆? In fact, presence of the three phases and of transitions between them may be
rationalized from analysis of the three limiting models which are combined in the Doniach
lattice solution we propose: a lattice-gas, a degenerate lattice-gas and Doniach's model.
Analysis of the phase diagrams of Fig. 7 in terms of the limiting models is given in the
Appendix.
14
1.0
0.8
t
0.6
0.4
0.2
(a)
Dis
Gas
Ord
8
.
0
2
.
1
1
1.2
1.0
0.8
t
0.6
0.4
0.2
Gas
(b)
Dis
0.7
0.6
0.5
0.4
Ord
−9 −8 −7 −6 −5
µ
−12 −9 −6 −3
µ
0
FIG. 7: Model phase diagrams depend on model parameters. In (a), ∆ is held fixed, while
K is varied (numbers represent values for K/J). In (b), K is held fixed, while ∆ is varied
(numbers represent values for ∆/J). In (a) we see that the gas-order line moves to higher
chemical potentials and the critical point disappears as K is lowered. (b) shows that
increasing ∆ dislocates both the gas-ordered chains coexistence line as well as the
order-disorder line, while the critical point disappears.
Finally, we would like to compare Doniach's lattice solution phase diagram with the phase
diagram of the original model, which was given in terms of pressure and temperature. From
thermodynamics, lateral pressure Π, conjugate to area A, is given by:
Π = −
1
A
Φ(T, A, µ, H) .
(24)
For our model, the thermodynamic grand-potential Φ is given by Eq. 21.
Figure 8 displays the lateral pressure Π versus temperature t phase diagram, for the
same model parameters as in Fig. 4. The gas phase is present at low pressure and higher
temperatures, as for usual fluids. At higher pressures, the two fluid phases are separated
by a coexistence line, with the ordered chain liquid (Ord) at lower temperatures, and the
disordered chains liquid (Dis) at higher temperatures. The coexistence pressure Π∗ between
the two liquid phases rises steeply as temperature is increased.
In the case of the model by Doniach, pressure is a linear function of temperature at
the order-disorder coexistence line, which ends at a critical point. In that case, a unique
order parameter is present, with m(a) (see Eq. 4) and the chain order-disorder transition
accompanies the pseudo -density transition.
In the model we propose, the chain order-
disorder transition is associated to a true density transition.
At low chemical potential and pressure, chain disordering is accompanied by a density
gap, which goes exponentially to zero at higher potentials. The linear dependence between
coexistence temperature and 'pressure' disappears.
15
b
c
b
c
b
c
b
c
b
c
b
c
b
c
b
c
b
c
b
c
0.5
0.4
0.3
Π
0.2
0.1
0.0
Ord
Dis
tc
Gas
0.8
0.9
t3
t0
0.7
t
0.5
0.6
FIG. 8: Lateral pressure vs versus temperature. Dashed line indicates the high density
limit of the order-disorder transition, temperature to.
V. PHYSICAL INTERPRETATION: MONOLAYERS VS BILAYERS
In order to interpret our findings in terms of the two systems of interest, monolayers
and bilayers, we must analyse the differences between inter-particle interactions in the two
systems, as well as the physical boundary conditions involved.
The anisotropic organization of the phospholipid molecules in layers is a consequence of
the fact that those are amphiphilic molecules, with a hydrophylic polar headgroup which
mixes with water and a hydrophobic hydrocarbonic tail which would phase separate were it
not attached to the polar headgroup. This nematic like structure is common to bilayers and
monolayers. However, there are specific aspects of the interactions between lipids and water
which make them different physical systems. While headgroups are in contact with water in
both systems, hydrocarbonic tails turn to the air subphase and have no contact with water,
independently of the distance between lipids, in the case of monolayers, whereas for bilayers
hydrophobic chains turn to the hydrophobic bilayer core, but water penetration increases as
lipid molecules go apart.
Our model system is a plane system in two dimensions. Monolayers are truly two-
dimensional, while bilayers may be seen as two weakly interacting spherical monolayers.
Monolayers reside in the interface between water and air, while bilayers are in bulk water,
albeit each leaflet of the bilayer could be taken as located on a water-hydrocarbon interface.
Monolayers may be manipulated both through direct compression, as well as through heat-
ing, implying a line of disordering transitions in the pressure-temperature plane. Bilayer
16
b
b
c
behaviour is probed through temperature variations only, and the disordering transition
occurs at a single temperature.
In the following subsections we analyse the differences pointed and out above and the
relation to our model.
A. Monolayers
A monolayer is constituted by lipid particles, whose chains suffer Van der Waals at-
traction. Lipid headgroups rest on the water surface, and lipid chains do not get in con-
tact with water molecules, which allows us to take null lipid-water interaction parameters,
ǫlip,w = ǫow = ǫdw = 0. On the other hand, water-water "bonds" are surface "bonds",
from now on labelled as ǫsurf
ww . When lipid molecules become dispersed, water molecules at-
tract strongly between themselves yielding large surface tension. At very low temperature
and very low pressure, one expects the model system to go into a 'gas' lipid phase, since
water-water interactions are dominant over lipid-lipid interactions.
Much of the experimental investigation of monolayers is given in terms of Langmuir
pressure-area isotherms, which present two coexistence plateaus, one between the gas and
the expanded liquid, at lower pressure, and the second one between the expanded and the
condensed liquid phases [1, 2], for which the discontinuity in area per molecule is an order of
magnitude lower. If compression is further increased, collapse of the monolayer comes about
[24]. Our model isotherms displayed in Fig. 9 compare well qualitatively to experimental
plots [25]. Both transitions are present, at different orders of magnitude both for pressure
and area per molecule. The lattice, of course, limits the minimum area, so that isotherms
increase steeply at the lower limit, differently from the experimental system, which, besides,
may be allowed to expand into a 3rd dimension. (see Figs. 5 an 6d-e)
A critical point for the ord-dis transition is absent for the parameters explored in this
study. However, it may be present for a different set of parameters, as explained in the
appendix.
17
2.0
1.6
−1
0.20
(a)
0.15
Π
0.10
(b)
80
60
40
20
0
−12
−16
−20
Π
1.2
0.8
0.4
0.0
−2
−3
−4
−5
−6
−7
−8
−10
−12
1.0
1.1
1.2
a
0.05
0
2
4
a
−40
6
8
FIG. 9: Pressure vs lipid area isotherms. ∆ = 0.6, K = 1 and Ω = 1000. Figures indicate
temperatures referred to the high density limit of the order-disorder transition, t0 (see Fig
9) as (t − t0) ∗ 1000. Isotherms present two plateaus corresponding to the order-disorder
and the gas-liquid transitions, for t3 < t < t0. Isotherms span a large range of areas, and
are thus represented separately for the two transitions. (a) On the left, pressures are
higher and area variation remains within 30% upon the order-disorder transition. The
order-disorder transition is present only below t0. (b) On the right, pressures are an order
of magnitude lower, and area may vary by a factor of 10, along the gas-liquid disordered
liquid above t3 (t0 − 12/1000) and gas-ordered liquid below t3.
B. Bilayers
In the case of a bilayer, the gas-liquid transition would correspond to membrane disag-
gregation and some critical micellar parameter [26], which is not of interest in the study of
biomembranes. As for the integral vesicle thermal phases, differently from the monolayer
case, water-lipid interactions are essential: they are the source of the "hydrophobic" inter-
action, ǫhydroph ≡ ǫbulk
ww − ǫlw, with ǫow 6= 0 and ǫdw 6= 0. Also, water-water "bonds" in the
bulk are "looser" than at the surface, and we thus label them as ǫbulk
ww .
In relation to pressure effect, bilayers may be considered to be in a tension-free state,
which corresponds [9, 10, 27] to a situation of zero lateral pressure Π. Different authors
propose an equivalence at some specific lateral pressures, but lately this equivalence has
been questioned.
18
b
c
b
c
b
c
b
b
c
b
c
b
c
C. Bilayers vs monolayers
How then are we to associate the model phase diagrams, Fig. 4 and Fig. 8, to physical
monolayers and bilayers?
A reasonable simplification is to take lipid-water interactions independent of lipid state,
with ǫow = ǫdw = ǫlw. This assumption yields the following relations between monolayer and
bilayer parameters (see Eqns. 8 - 11):
J M = J B ,
∆M = ∆B ,
K M = K B + (ǫsurf
ww − ǫbulk
ww ) + 2ǫlw ,
and
I M = I B + 4(ǫsurf
ww − ǫbulk
ww ) + 4ǫlw .
(25)
(26)
(27)
(28)
What are the implications of the difference in energy parameters for the two systems?
Let us first analyse the differences between parameters K M and K B. Interactions between
implies ǫsurf
water molecules on the surface are more stable than between molecules in the bulk, which
ww . Thus we have K M > K B. This is an important result, since it implies
that data for the two systems cannot be mapped onto the same phase diagram. We inspect
ww > ǫbulk
Fig. 7a taking into account this point. Variation of K has two simultaneous effects. As
it increases, (i) it displaces the gas-liquid line to lower chemical potential, as should be
expected, since it favours lipid-lipid interactions; (ii) it turns the order-disorder coexistence
line near the triple point less dependent on temperature.
The last effect has implications on the area discontinuity upon the order-disorder transi-
tion. If one focuses on this transition at some specific temperature, the discontinuity in area
is smaller for larger K. This means that, if effective chemical potential µeff is the same for
both systems, adopting the same transition temperature implies obtaining different areas
for the two systems at coexistence. Interestingly, this is what happens with the experimen-
tal systems: if one looks for the equivalence at the same transition temperature, areas are
different. If, on the other hand, equivalence is sought for from the same area gap, the two
transitions are found at different temperatures.
19
A further difference between the two systems arises if we try inspection of the pressure-
temperature phase diagram (Fig. 8). Different suggestions can be found in the literature for
the lateral pressure on monolayers which would render them equivalent to bilayers [10, 27].
If, for the sake of further analysing the pursuit of equivalence between the two systems, we
ignore the differences in K, it is possible to establish a relation between the lateral pressures
of the two systems. We consider a particular thermodynamic state for the model system,
which corresponds to a point in the µ vs t phase diagram, and to the same order parameter
values for m and n (Eqns. 23). Note that if the two systems, monolayer and bilayer are
considered to be in the same thermodynamic system, mB = mM and nB = nM, besides
having equal K parameters, both systems must have equal effective chemical potentials µ,
i.e. µM = µB. What are the implications on lateral pressure?
Lateral pressure is related to the grand-potential through Eq. 23. From the definition of
the grand potential [28],
Φ = hEi − T S − µlNl − µwNw)
= hEint − T S − µNl − (µw + 2ǫww)Ai
(29)
For the same thermodynamic state, for K M = K B, the grand-potential for the two
systems will differ only through the constant terms, since density of lipids and entropy must
be the same. Thus
ΦB − ΦM = −(cid:0)K B − K M(cid:1) Nll − 2(cid:0)ǫbulk
ww − ǫsurf
ww(cid:1) A ,
(30)
which yields the following simple relation for the lateral pressures of the two systems:
ΠM − ΠB = 2(cid:0)ǫsurf
ww (cid:1) ,
ww − ǫbulk
(31)
since Φ = −ΠA.
For the null pressure of the bilayer, corresponds a positive lateral pressure on the mono-
layer,
since ǫsurf
ww > ǫbulk
ww .
ΠM = 2(ǫsurf
ww − ǫbulk
ww ) ,
(32)
This result is qualitatively in agreement with several proposals of the literature based on
experimental measurements [8 -- 10] and gives it an interpretation in terms of the statistical
20
3.0
2.5
a
2.0
1.5
1.0
0.6
1
.
0
5
0 . 0
0
.
0
0.7
t
0.8
FIG. 10: Area per lipid as function of temperature at fixed pressure for the order-disorder
transition. K/J = 1 and ∆/J = 0.6. ΠM = 0.1 or ΠB = 0 (see text) for K M = K B and
values of ǫlw/J
are as indicated. Density gap decreases as mixing with water is favoured.
model. In particular, it is also in line with a phenomenological analysis proposed by Marsh
[10], in which the monolayer pressure corresponding to the membrane thermodynamic state
at the main transition would be numerically equal to the hydrophobic free energy density.
The origin of the difference in pressures would be the water surface tension, a consequence
of "stronger" hydrogen bonds on the surface, as compared to bulk water.
Figure 10 illustrates the behaviour of the density gap for different ratios of the lipid-
water interaction to water surface tension, at fixed pressure, under the artificial condition
of equivalence (K M = K B). Increasing lipid-water interaction, with respect to water surface
tension, transition temperature is decreased, while the area discontinuity increases.
However, the result for the equivalence pressure just presented, as stated before, is based
on the artificial assumption of equal interaction coefficient K for the two systems. But Eq. 27
shows that this assumption is inconsistent with the presence of either water-water or water-
lipid interactions. Therefore, the difference in the interaction strength of water molecules on
the surface and in bulk, as well as the presence of water-lipid interactions only for the bulk
lipid layer make inconsistent the hypothesis of equivalence between monolayers and bilayers,
and explains the difficulty in aligning simultaneously both the transition temperature and
the density gap [27].
21
VI. FINAL COMMENTS
We have proposed a generalization of the two-state model for lipid layers, which allows
an exact description of local density. This is essential for the investigation of the effect of
charges, in the case of dissociating lipids.
Inspection of model properties with respect to the relation between density and chain
order led to some additional conclusions:
(i) for monolayers, the model describes both the liquid transitions (order-disorder or
condensed liquid-expanded liquid), in good qualitative agreement with experimental studies;
(ii) analysis in terms of the different model interactions between lipids and water for
monolayers and bilayers yields an explanation for the difficulty in establishing the equivalence
between the two experimental systems.
Further investigation of the model system in the presence of charges, both for dissociating
headgroups as well as for dipolar headgroups, is underway.
VII. ACKNOWLEDGMENTS
We thank Eduardo Henriques at UFPEL for pointing out the possibility of adapting
our model to the study of monolayers and to Mario Tamashiro at UNICAMP for many
conversations on the theme of our work.
APPENDIX: LIMITING MODELS
Our model may be thought as a composition of three models: (i) the order-disorder
Doniach model, of density one, (ii) a simple lattice gas, and (iii) a degenerate lattice gas.
The three limiting models are obtained if one of the three values for site variables σ is
discarded. Model (i) results from making site variables σ equal to 1 and −1. Model (ii)
results from restricting site variables σ to 0 and +1. Model (iii) is obtained if site variables
σ are taken as 0 and −1. Under such restrictions, each one of the three limiting models may
be mapped on the two-state Ising model, given by
EIsing = − JX(ij)
sisj − HXi
si ,
(33)
22
with particle states s assuming two possible values, s = +1 or s = −1. Model parameters
J and H are different for each model, and if given in terms of the Doniach lattice gas
parameters (Eqs. 8 9 10 ), are as follows: For model (i),
For model (ii)
Ji = J ;
Hi(T ) = 4∆ −
1
2β
ln Ω .
Jii =
1
4
(J + 2∆ + K) ;
Hii(µ) = J + 2∆ + K +
1
2
µ .
Finally, for model (iii), we have
Jiii =
1
4
(J − 2∆ + K) ;
and
Hiii(T, µ) = J + K − 2∆ +
1
2
µ +
1
2β
ln Ω .
(34)
(35)
(36)
(37)
(38)
(39)
The phase behaviour of each of the three models may be obtained by adapting well-known
results for the Ising model for ferromagnetism. The Ising model presents a coexistence line
at H = 0 which ends at a critical temperature TC, given by
kBTC
J
= 4 .
(40)
As a result, a coexistence line and a critical point exists for each of the three limiting
models. Model (i) displays a coexistence line at fixed temperature
t0 =
8∆
J ln Ω
(41)
between the disordered (m = 1) and ordered chain (m = −1) liquids. Model (ii) presents
coexistence at fixed chemical potential
µ0 = −2(J + 2∆ + K) ,
23
(42)
1.2
1.0
0.8
t
0.6
0.4
0.2
Gas
Dis
0.7
0.6
0.5
0.4
Ord
−12
−9
−6
µ
−3
0
FIG. 11: Coexistence lines for DLG (symbols) and coexistence lines for the limiting
models: Doniach's model (horizontal dashed lines), lattice gas (vertical dashed lines) and
degenerate lattice gas (sloped dashed lines). Critical points for the limiting models are
seen only for the case of the degenerate lattice gas. For the other two limiting systems,
critical points are outside frame. For ∆ = 0.7, the DLG presents no Gas-Dis line. Also, the
critical temperature Gas-Liquid line of the limiting degenerate lattice gas model is below
the temperature of the Ord-Dis of the limiting Doniach model.
between a gas (n = 0) and a simple liquid (n > 0), with a critical point at tC,(ii) = 1 +
2∆/J + K/J. Finally, in model (iii) a gas (n = 0) and a degenerate liquid (n > 0) coexist
at µ(t) = −2(J − 2∆ + K) − kBT ln Ω, with a critical point at
tC,(ii) = 1 − 2
∆
J
+
K
J
.
(43)
In Fig. 11 we compare phase coexistence lines of previous Fig. 7b with coexistence lines
for the limiting models (i)-(iii), at different values of ∆. As can be seen, for the case in
which the critical temperature of the limiting model (ii) is lower than the temperature for
the chain order-disorder transition the line G-Dis disappears, as expected. Similar analysis
explain previous Fig. 7 a. The analysis of the limiting models also allows us to expect a
critical point at the end of the Ord-Dis coexistence line if t0 > 4.
[1] V. M. Kaganer, H. Mohwald, and P. Dutta, Rev. Mod. Phys. 71, 779 (1999).
[2] H. Mohwald, in Structure and Dynamics of Membranes, Vol. 1, edited by R. Lipowsky and
E. Sackmann (North-Holland, 1995) Chap. Handbook of Biological Physics, pp. 161 -- 211.
24
b
c
b
c
b
c
b
c
b
c
b
c
[3] M. Bloom, E. Evans, and O. G. Mouritsen, Quarterly reviews of biophysics 24, 293 (1991).
[4] S. Tristram-Nagle and J. F. Nagle, Chemistry and Physics of Lipids 127, 3 (2004).
[5] J. F. Nagle, The Journal of Chemical Physics 58, 252 (1973).
[6] J. F. Nagle, The Journal of Chemical Physics 63, 1255 (1975).
[7] S. Marcelja, Biochimica et Biophysica Acta (BBA) - Biomembranes 367, 165 (1974).
[8] J. F. Nagle, The Journal of Membrane Biology 27, 233 (1976).
[9] D. W. R. Gruen and J. Wolfe, Biochimica et Biophysica Acta (BBA) - Biomembranes 688,
572 (1982).
[10] D. Marsh, Biochimica et Biophysica Acta (BBA) - Reviews on Biomembranes 1286, 183
(1996).
[11] Y. Ni, S. M. Gruenbaum, and J. L. Skinner, Proceedings of the National Academy of Sciences
110, 1992 (2013).
[12] A. Caille, A. Rapini, M. J. Zuckermann, A. Cros, and S. Doniach, Canadian Journal of
Physics 56, 348 (1978).
[13] S. Doniach, The Journal of Chemical Physics 68, 4912 (1978).
[14] O. G. Mouritsen, A. Boothroyd, R. Harris, N. Jan, T. Lookman, L. MacDonald, D. A. Pink,
and M. J. Zuckermann, The Journal of Chemical Physics 79, 2027 (1983).
[15] J.-F. Baret and J.-L. Firpo, Journal of Colloid and Interface Science 94, 487 (1983).
[16] D. A. Pink and D. Chapman, Proceedings of the National Academy of Sciences 76, 1542
(1979).
[17] T. Heimburg, Thermal Biophysics of Membranes (Tutorials in Biophysics), 1st ed. (Wiley-
VCH, 2007).
[18] P. F. Almeida, Biophysical journal 100, 420 (2011).
[19] M. N. Tamashiro, C. Barbetta, R. Germano, and V. B. Henriques, Phys. Rev. E 84, 031909
(2011).
[20] M. T. Lamy-Freund and K. A. Riske, Chemistry and Physics of Lipids 122, 19 (2003).
[21] R. P. Barroso, K. A. Riske, V. B. Henriques, and M. T. Lamy, Langmuir 26, 13805 (2010).
[22] V. B. Henriques, R. Germano, M. T. Lamy, and M. N. Tamashiro, Langmuir 27, 13130
(2011).
[23] C. E. I. Carneiro, V. B. Henriques, and S. R. Salinas, Physica A: Statistical Mechanics and
its Applications 162, 88 (1989).
25
[24] K. Y. C. Lee, Annual Review of Physical Chemistry 59, 771 (2008).
[25] S. Marcelja and J. Wolfe, Biochimica et Biophysica Acta (BBA) - Biomembranes 557, 24
(1979).
[26] C. S. Shida and V. B. Henriques, International Journal of Modern Physics C 09, 801 (1998).
[27] J. F. Nagle and S. Tristram-Nagle, Biochimica et Biophysica Acta (BBA) - Reviews on
Biomembranes 1469, 159 (2000).
[28] H. B. Callen, Thermodynamics and an Introduction to Thermostatistics, 2nd ed. (Wiley, 1985).
26
|
1704.08313 | 1 | 1704 | 2017-04-26T19:35:13 | Drugs and Drug Delivery Systems Targeting Amyloid-\b{eta} in Alzheimers Disease | [
"physics.bio-ph"
] | Alzheimer's disease (AD) is a devastating neurodegenerative disorder with no cure and limited treatment solutions that are unable to target any of the suspected causes. Increasing evidence suggests that one of the causes of neurodegeneration is the overproduction of amyloid beta (A\b{eta}) and the inability of A\b{eta} peptides to be cleared from the brain, resulting in self-aggregation to form toxic oligomers, fibrils and plaques. One of the potential treatment options is to target A\b{eta} and prevent self-aggregation to allow for a natural clearing of the brain. In this paper, we review the drugs and drug delivery systems that target A\b{eta} in relation to Alzheimer's disease. Many attempts have been made to use anti-A\b{eta} targeting molecules capable of targeting A\b{eta} (with much success in vitro and in vivo animal models), but the major obstacle to this technique is the challenge posed by the blood brain barrier (BBB). This highly selective barrier protects the brain from toxic molecules and pathogens and prevents the delivery of most drugs. Therefore novel A\b{eta} aggregation inhibitor drugs will require well thought-out drug delivery systems to deliver sufficient concentrations to the brain. | physics.bio-ph | physics | Volume 2, Issue 3, 332-358.
DOI: 10.3934/molsci.2015.3.332
Received date 16 May 2015,
Accepted date 20 July 2015,
Published date 29 July 2015
http://www.aimspress.com/
Review
Drugs and drug delivery systems targeting amyloid-β in Alzheimer's
disease
Morgan Robinson 1, Brenda Yasie Lee 1 and Zoya Leonenko 1, 2, 3, *
1 Department of Biology, University of Waterloo, Waterloo, Ontario, Canada
2 Department of Physics & Astronomy, University of Waterloo, Waterloo, Ontario, Canada
3 Waterloo Institute for Nanotechnology, Waterloo, Ontario, Canada
* Correspondence: Email: [email protected]; Tel: +1 519-888-4567 ext. 38273.
Abstract: Alzheimer's disease (AD) is a devastating neurodegenerative disorder with no cure and
limited treatment solutions that are unable to target any of the suspected causes. Increasing evidence
suggests that one of the causes of neurodegeneration is the overproduction of amyloid beta (Aβ) and
the inability of Aβ peptides to be cleared from the brain, resulting in self-aggregation to form toxic
oligomers, fibrils and plaques. One of the potential treatment options is to target Aβ and prevent
self-aggregation to allow for a natural clearing of the brain. In this paper, we review the drugs and
drug delivery systems that target Aβ in relation to Alzheimer's disease. Many attempts have been
made to use anti-Aβ targeting molecules capable of targeting Aβ (with much success in vitro and in
vivo animal models), but the major obstacle to this technique is the challenge posed by the blood
brain barrier (BBB). This highly selective barrier protects the brain from toxic molecules and
pathogens and prevents the delivery of most drugs. Therefore novel Aβ aggregation inhibitor drugs
will require well thought-out drug delivery systems to deliver sufficient concentrations to the brain.
Keywords: amyloid; amyloid aggregation; inhibitor drugs; Alzheimer's disease; drug delivery
systems; blood brain barrier; nanotechnology
Abbreviations: Alzheimer's disease (AD), amyloid-beta (Aβ), blood-brain barrier (BBB), amyloid
precursor protein (APP), polyethylene glycol (PEG), monoclonal antibodies (MAb), beta-secretase 1
(BACE1), passive immunotherapy (PI), atomic force microscopy (AFM), retro-inverso (RI),
molecular dynamics (MD), cell-penetrating peptide (CPP), polylactic-co-glycolic acid (PLGA),
nanoparticle (NP), tight junctions (TJ), transendothelial electrical resistance (TEER), P-glycoprotein
(P-gp), advanced glycation end products (RAGE), transferrin receptor (TfR), diphtheria toxin
333
receptor (DTR), receptor-mediated transcytosis (RMT), low density lipoprotein related protein
(LRP1), apolipoprotein E (ApoE) rabies virus glycoprotein (RVG), gold nanoparticle (AuNP), poly
lactic acid (PLA), absorptive-mediated transcytosis (AMT), transactivator of transcription (TAT),
Madin-Darby canine kidney (MDCK)
1. Introduction
With people living longer than ever and aging populations increasing around the world, the
health concerns associated with age will continue to increase. Those afflicted with Alzheimer's
disease represent a very large portion of the geriatric population. Accounting for 50‒80% of cases of
dementia, AD is the leading cause of dementia worldwide, with over 36 million registered cases in
2009. It is expected that the incidence of AD will double every twenty years, with approximately 66
million worldwide by 2030 and 115 million cases globally by 2050 [1]. With no cure available,
specific treatment options for AD are limited to treating the symptoms of dementia, not the root
cause. Five FDA approved drugs are available for improving memory and cognitive function in AD
patients. These drugs attempt to increase the amount of neurotransmitters in the brain, which have
been shown to improve cognition in mild to moderate AD patients for 6 months [2]. Not only are the
treatment methods non-specific to AD but a sensitive and specific diagnostic test for AD is still
unavailable. The only conclusive diagnosis for AD is performed during a postmortem autopsy on the
brain, where plaques can be directly visualized.
Early detection of AD is near but not yet reliable and preventative treatments in the
pre-dementia phase (called prodromal AD) are not available as it has only recently been defined.
Psychiatric evaluations involving memory tests are the standard diagnostic tool with neural imaging
used as a supplemental diagnostic and is a new and exciting emerging technology which will be
useful for diagnosing prodromal AD [3,4]. Lack of detection is especially concerning as evidence
suggests symptoms may require a decade to manifest after the onset of pathophysiology in AD [5].
The amyloid beta (Aβ) hypothesis
There are several theories which have been proposed to explain the progression of Alzheimer's
disease with the amyloid cascade hypothesis remaining the most widely accepted. This hypothesis
states that the primary neurologic insult is caused by toxic and soluble Aβ oligomers. The
trans-membrane amyloid precursor protein (APP) is cleaved by β and γ secretases to produce Aβ
monomers [4], which misfold to form toxic β-sheet oligomers and eventually form larger fibrils and
plaques. After the enzymatic processing of APP, Aβ monomers are free to misfold back onto
themselves through side chain interactions producing a hairpin structure with residues in the central
region and N-terminus forming intermolecular hydrogen bonds with other monomers [6]. Binding
sites for monomers always appear on the outside of the growing oligomer, turning into nucleation
sites for further growth into toxic oligomers and eventually fibrils and plaques.
Literature in the last decade has implicated small oligomers as the most toxic of amyloid
aggregate species, with plaques and fibrils being non-toxic even though they may still be a source of
free amyloid [7]. Current reports suggest that non-specific interactions of toxic soluble amyloid
oligomers with the cell membrane play an important role in the mechanism of cytotoxicity [7,8].
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
334
Recent studies in amyloid lipid membrane interactions have been able to show that amyloid
oligomers aggregate on the surface of the cell membrane, deforming it [7-9]. In extreme cases,
trans-membrane channels [10-12] and pores are formed that disrupt membrane integrity [13,14].
Effects on the cell membrane caused by amyloid may be more subtle, including: alterations in
synaptic plasticity, receptor protein distribution and modification of signaling pathways prior to the
occurrence of more severe deformations [8,13]. Aβ molecules are found in all humans, regardless of
age or disease but the natural role of these amyloid are largely unknown. It has been suggested that
the Aβ monomer may play a role in important signaling pathways in the brain [15] and is likely to
have neuroprotective properties at low concentrations [16]. It appears that healthy individuals are not
susceptible to amyloid induced cytotoxicity and can clear amyloid from the brain before it reaches
neurotoxic levels by balancing amyloid production and clearance [17]. If amyloid burden can be
reduced, it is speculated that it may be possible to slow the progression of Alzheimer's disease.
2. Therapeutic strategies in AD targeting amyloid aggregation pathways
Insight into amyloid aggregation has led to various approaches in the last two decades to slow
or prevent amyloid aggregation and improve clearance from the brain. One approach is to prevent
aggregation by using molecules/ligands which directly bind to and modify or inhibit aggregation of
amyloid. These include: PEG nanoparticles [18], lipid based nanoparticles containing phosphatidic
acid and cardiolipin [19], small molecules (like curcumin, melatonin) [20-24], monoclonal antibodies
(MAbs) directed against Aβ [25-30] and various other peptidic aggregation inhibitors [31-35]. Another
approach that has been proposed with favorable results is active immunotherapy or Aβ vaccines; this
was found to lessen amyloid burden on the brain in mouse models [36]. However, concerns with
safety have limited the use of active amyloid vaccines. In one clinical trial, 6% of patients developed
meningoencephalitis, a dangerous inflammation of the meninges and the brain [37]. Antibodies
against amyloid, synthetic or native, signal microglial cells to facilitate uptake and autophagy of
amyloid, clearing it from the brain. It is also possible to target other clearance mechanisms,
upregulating proteins responsible for moving amyloid out of the brain, via receptor mediated BBB
efflux or enzymatic degradation pathways [38]. A third approach is to prevent aggregation at the
source by targeting its precursor, either APP itself or the enzymes which cleave it. The effects of
APP and secretase inhibition is not well established and dangerous consequences on downstream
pathways are suspected as they play roles in many neural processes [39,40].
2.1. APP cleavage by secretases
Reduction in amyloid burden can be achieved by targeting the enzymatic pathway responsible
for cleaving APP into Aβ monomers. The rate limiting enzymatic step involves beta-secretase 1
(BACE1) cleavage of APP and presents a key target for inhibition [39]. Secretase inhibitor strategies
include RNA interference to prevent translation of the enzyme, small molecules to reduce the
enzyme's activity and MAb to clear the enzyme [41-44]. Lentiviral vectors expressing siRNAs
directed against BACE1 transcripts were shown to reduce amyloid production and cognitive deficit
in APP transgenic mouse models [43]. Small molecule inhibitors of secretases are non-specific and
have displayed unpredictable off target effects that have led to cancellation of some clinical trials [44],
while larger more specific therapeutics show limited BBB permeation [41]. Antibodies to BACE1
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
335
were discovered to specifically target BACE1 and prevent amyloidosis in human cell lines, primary
neurons, and in vivo mouse and non-human primate models, but BBB permeation still represents a
large barrier to their effectiveness [41]. Although the knockout of BACE1 in mouse studies yielded
insignificant consequences [45], the effects on BACE1 substrates (such as myelination, retinal
homeostasis, synaptic function and brain circuitry) are still not well understood and more studies are
required. Combination therapies represent an exciting opportunity to achieve improved amyloid
reduction in the brain but have not been explored in great detail. Recently, anti-Aβ monoclonal
antibody Gantenerumab directed at amyloid and a small molecule BACE1 inhibitor were assessed in
a London mouse model of AD [46]. Modulators and inhibitors of gamma secretases have also been
explored, complete inhibition is not feasible due to the role it plays in critical Notch signaling
pathways, while modulators are able to preferential target gamma secretase activity on APP alone
and are of great therapeutic interest [40]. With at least 10 small molecule BACE1 inhibitors in
current clinical trials, BACE1 represents a key target for therapeutic intervention of amyloid
targeting intervention.
2.2. Passive immunotherapy targeting amyloid
Passive Immunotherapy (PI) uses MAbs with high-specificity to the various species of Aβ to
reduce aggregation and promote immunogenic clearance from the brain [28,29]. Three non-mutually
exclusive pathways for the PI mechanism have been suggested and may all act in parallel [47].
MAbs are expected to directly trigger an immune response against Aβ deposits in the brain,
increasing microglial uptake [29]. MAbs can also prevent aggregation of and disaggregate fibrils and
plaques through competitive processes; this is supported by early AFM studies on MAbs like
m266.2 [27]. MAbs have been suggested to improve cognition and amyloid burden in mouse models
through a proposed "sink" effect where the reduction of free soluble amyloid in the blood causes a
shift in amyloid equilibrium to favor efflux from the brain [26]. MAbs do not cross the BBB
efficiently, reaching a maximum of 0.11% at 1 hour after injection, which adds weight to the sink
hypothesis [48]. In AD, BBB integrity is compromised and therefore increases the risks associated
with over-activating microglial cells and pro-inflammatory response. Various dangerous risks
including cereberal amyloid angiopathy and other conditions have been reported [49-52]. Attempts
to engineer safer MAbs, with reduced effector response and improved BBB permeation, are
currently being explored [29]. Another obstacle in using PI is that while largely successful in many
preclinical mouse models, both at reducing amyloid burden and improving cognition [25,26,30,52],
MAb therapies for use in humans has been unsuccessful at improving the key markers of AD
(cognition, memory and learning) in clinical trials [53-56].
Solanezumab (Sol) is the humanized IgG1 MAb m266.2, which recognizes the central region of
Aβ13‒28 and binds soluble amyloid species. Solanezumab was shown to have no significant effect on
improving the primary outcomes of patients in two phase 3 clinical trials, though no adverse events
could be associated with it [53]. A more detailed analysis yielded mildly encouraging results for a
subgroup of mild AD patients [53]. This trial is ongoing while other preventative trials of
Solanezumab have begun. Bapineuzumab (which recognizes N-terminal amyloid and binds to fibrils
and plaques) was shown to have no significant effect on improving cognition of patients in two phase
3 clinical trials. Coupled with a safety concern, where MRI imaging showed abnormalities associated
with vasogenic edema, clinical trials were discontinued [55]. These setbacks have led experts to a
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
336
shift in expectations for amyloid intervention therapies and suggest that PI may not be effective after
the onset of symptoms due to the significant amount neuron loss [54,56]. To test this premise,
clinical trials have shifted to preventative studies, with the newcomer Crenezumab launching a
five-year preventative trial in 2012 [57]. Crenezumab was engineered with an IGg4 backbone which
causes a milder immune response yet still binds with high affinity to amyloid monomers, oligomers
and fibrils; it was shown to be neuroprotective and increased uptake of Aβ by microglial cells in
mouse models [25,57]. Another PI called Gantenerumab, which binds both the N-terminus and
mid-region of Aβ but predominantly fibrillar species, was being tested in phase 3 trials with
prodromal (pre-dementia) and mild AD patients [58]. However these studies have been cancelled due
to a lack of efficacy, findings from the trial have not yet been published. Since amyloid insult may
occur as early as 20 years prior to the onset of symptoms [5], early diagnostic testing stands as an
important obstacle to effective treatments. As paradigms shift to preventative strategies, engineered
MAbs and other peptide inhibitors of amyloidosis represent potentially safer and more effective
alternatives.
2.3. Peptide inhibitors of amyloid aggregation
In 1996, Tjernberg et al. reported the use of amyloid peptide fragment KLVFF as an aggregation
inhibitor and showed proof of principle for the use of peptide-based ligands built from the sequence
of amyloid itself [35]. Although aggregation is still seen, a noticeable decrease in fibrillization
suggests that these peptides may serve as ligands with the ability to affect the dynamics of fibril
formation. This study systematically tested 31 decamer Aβ fragments, selected the decapeptides with
highest affinity, truncated and mutated them to determine the minimum sequence needed to prevent
binding, and finally identified the pentapeptide KLVFF (Aβ16-20) [35]. They found that the sequence
KLXXF was critical for amyloid binding. The same year, Ganta et al. showed how Aβ15-25 attached
to a disrupter element (repeated oligolysine) could reduce amyloid beta toxicity. Although this
peptide inhibitor did not block beta sheet interactions and fibril formation, it did cause changes in
aggregation kinetics and higher order structural changes in fibrils by shortening the length of the
fibrils while increasing the amount of fibril entanglement [32]. This highlights the importance of
how the aggregated structure of A can affect cytotoxicity. Pallitto et al. modified the recognition
sequence KLVFF by adding repeating oligoproline units. The ring structure of the proline side chain
disrupts beta sheet interactions, limiting the stacking ability of amyloid to grow into larger fibrils
through a dynamic competitive process [34]. They also found that shuffling the KLVFF amino acid
sequence still inhibited Aβ fibril formation and had nearly identical binding characteristics. This
implies that the overall hydrophobicity of the amyloid ligand is important for efficient binding with
amyloid [34]. These early preliminary studies have paved the way for the advanced design of next
generation peptide inhibitors.
With a foundation of knowledge built from early amyloid studies, it is possible to design
effective amyloid aggregation inhibiting drugs with various compositions for stability, binding
affinity, cell membrane and BBB permeability, and immune system evasion. Small peptides under
nine amino acids made with synthetic amino acid residues-N-methylated, dextrorotary
(D)-improves immune system evasion, proteolytic stability, and also interactions with the target
amyloid [59,60]. A peptide inhibitor, called OR2, was designed from the KLVFF sequence, modified
with a glycine spacer and charged amino acid residue at each terminus. These charged residues
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
337
improve aqueous solubility and disrupt fibril formation. OR2 was shown to modify early aggregation
of Aβ and protect SHSY-5Y cells from Aβ cytotoxicity [61]. Later, to improve proteolytic stability
and reduce immune response they substituted various amino acids with their corresponding
D-enantiomer. In order to maintain the biological activity of the original peptide a simple swap will
not due, reversal of the peptide bond is also required. This "retro inverso" version of the peptide,
newly designated RI-OR2, was shown to be effective at inhibiting oligomerization and improving the
survival of SH-SY5Y cells against Aβ toxicity, while also remaining stable in human blood serum
and brain extract for at least 24 hrs [62].
The design and screening of potential drugs and protein therapeutics which bind to and prevent
oligomerization using Computer Aided Drug Design (CADD) is a rapid and cost-effective technique
for developing and screening drug candidates [63]. Procedurally, the amyloid oligomerization
inhibitor drug design begins with a basic lipophilic amyloid recognition sequence: KLVFF, LVFFAE
and KVLFFAE, as identified in earlier studies [35]. Several amino acids in the sequence are
N-methylated to help prevent the inhibitor from contributing to the growth of Aβ oligomers, also
N-methylation may improve membrane permeability [64]. Next, various modifications and
substitutions to the peptide are made, such as additions of γ-diaminobutyric acid as an N-terminal
residue for improved interactions with amino acid D23 and substitution of lysine with ornithine, a
synthetic amino acid which improves electrostatic side chain interactions with E22 [63]. Other
improvements to peptide inhibitors can be made by substituting various lipophilic residues, which
may optimize hydrophobic interactions between the drug and amyloid target. Alternatively, one can
substitute lipophilic aromatic amino acids which have been shown to be important for peptide and
protein recognition including amyloid [60]. With this series of inhibitors (labeled the SG series),
leading candidates based on MD simulations were tested for their ability to block aggregation using
thioflavin T fluorescence assay, western blot and circular dichroism [63]. Strong correlations
between success in MD simulation and thioflavin T fluorescence assays were found, highlighting the
success of in silica drug design and screening. Furthermore, it has been confirmed in our lab group
using atomic force spectroscopy that SG inhibitors are able to reduce binding events between single
amyloid monomers [33]. Currently, various SG inhibitors with different modifications, D-amino acid
incorporation and various predicted binding orientations are being screened and tested both through
direct force measurements and in vitro assays.
Peptide based inhibitors present an alternative and appealing preventative strategy to MAb
therapies, as they are not as costly to produce, are smaller in size, versatile and intrinsically safer.
Antibody fragments and Fc engineered MAbs, which are designed to be safer alternatives to
traditional PI, do not offer significant advantages over peptide based inhibitors. Peptide inhibitors are
also easily modified for superior BBB permeation through the addition of targeting ligands and
shuttling molecules. Parthsarathy et al. modified a peptide inhibitor with a cell-penetrating peptide
(CPP) derived from an HIV regulatory protein. This was shown to improve delivery of the peptide to
cells and the brain, showing improvement in a transgenic mouse model [65]. Through a more
complicated process, MAbs can also be improved by making them bi-specific, able to bind amyloid
and some feature of the BBB for improved brain delivery [29,66]. Immunogenic clearance of Aβ has
been the focus of research, however as paradigms in AD treatment shift to prevention, inhibition of
Aβ aggregation may be suitable to allow natural clearing, to which peptide inhibitors have the
advantage.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
338
2.4. Nanoparticles targeting amyloid
Designing efficient drugs for targeting and interrupting early amyloid self-aggregation is
incredible challenging due to the small interaction surface area and lack of higher order structure of
amyloid monomers for binding. One suggested way to overcome this is through the use of various
nanoparticles (NP). PEGylated long circulating polylactic co-glycolic acid (PLGA) NPs have been
shown to capture monomeric and oligomer Aβ in solution and serum, the authors propose that a
possible "sink" mechanism may result if used for treating AD [18]. In a similar fashion nanoparticles
functionalized with the Aβ ligand, LVFFARK (Figure 1), were shown to protect SH SY5Y cells from
amyloid toxicity compared to the peptide alone, which although inhibited fibrillization, also had
exhibited strong self-assembly characteristics causing high cytotoxicity. These nanoparticles could
potentially act as peripheral sinks, like MAbs and the above mentioned PEGylated NP [67]. As it has
been suggested that anionic lipids in the cell membranes cause nucleation sites for amyloid
oligomerization, phosphatidic and cardiolipin lipid nanoliposomes were developed which interact
with amyloid oligomers and fibrils in solution and serum and could be used to target Aβ [19]. In a
more recent study curcumin incorporated into the nanoliposomes were compared to lipid ligand
nanoliposomes and shown to be even more effective at inhibiting oligomer and fibril formation [68].
These early molecular studies have led to further progress in delivery and targeting of amyloid which
will be discussed later in section 3.
Figure 1. PLGA nanoparticles modified with Aβ recognition peptide LVFFARK.
Shown as green on the cartoon above. Reprinted with permission from Xiong N, Dong
X-Y, Zheng J, et al. (2015) Design of LVFFARK and LVFFARK-functionalized
nanoparticles for inhibiting amyloid β-protein fibrillation and cytotoxicity. ACS applied
materials & interfaces 7: 5650-5662. Copyright © 2015 American Chemical Society.
All of the previous therapeutic strategies, though in principle effective at targeting various
aspects of the amyloid cascade, have yet to yield definitive disease-modifying results in human
clinical trials causing many to question the amyloid cascade hypothesis [54,56,69,70]. It is likely that
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
339
amyloid intervention is required prior to the formation of toxic amyloid oligomers [54,56], or that
amyloid is not the main cause of neural degradation but a byproduct that causes secondary
degeneration [69,70]. To verify the amyloid cascade hypothesis, anti-amyloidosis targeting strategies
need to be improved and tested in a relevant population. There are two major challenges that need to
be addressed before amyloid targeted therapies can be verified: (1) early diagnostic solutions need to
be found so preventative measures can truly be tested, prior to symptom onset and (2) strategies to
overcome the restrictive BBB and improve drug and diagnostic probe delivery to the brain are
necessary.
3. Drug delivery vehicles and pathways
Aside from the use of NPs as therapeutics against amyloid in AD, nanotechnology has received
significant attention for its efficiency through improved surface area in countless other applications.
For biomedical purposes and especially delivery across the BBB, there are some necessary and ideal
criteria that should be met for safety and efficiency; they should be non-toxic, non-immunogenic,
biodegradable, biocompatible, have prolonged blood circulation time, colloidal stability, surface
enhancement for BBB permeation, controlled drug release profile and size. The potential efficacy for
the delivery of therapeutics across the BBB for the treatment of various diseases using nanoparticles
has been well documented using many variations of the same basic structure. These NP drug delivery
systems use a basic nanocore structure, such as amphiphilic, polymeric, metallic or hybrid structures.
The surface of the NP is important for efficiency and safety, it must be suitably modified for
recruitment across the BBB, and finally it will contain the drug target either as a ligand on the
surface of, or contained within the NP.
3.1. The blood brain barrier (BBB)
The blood brain barrier is a highly selective, regulated and efficient barrier that protects the
brain from unwanted molecules and pathogens. It is also the single largest hurdle that needs to be
overcome in order to deliver potential therapeutic and diagnostic agents to the brain. The lack of
therapeutics for the treatment of Alzheimer's and other brain diseases is not only due to the lack of
effective drugs, but due to their inability to cross into the brain from the blood through the
endothelium [71]. In the human brain there is approximately 100 billion capillaries and a BBB
surface area of 20 m2, as compared to 0.021 m2 for the blood-cerebral spinal fluid barrier [72].
Therefore, most of the entry into the brain is controlled by the BBB, an interface separated by brain
endothelial cells on the blood side and astrocytes and pericytes on the brain side [73]. The BBB is
comprised of endothelial cells "glued" together with multiple binding proteins (occludins, claudins
and junctional adherin molecules) to form tight junctions (TJs) and adherin junctions (AJs). A large
portion of brain homeostasis is regulated by influx and efflux at the BBB through these junctions.
The BBB has the abilities to prevent entry of and actively remove unwanted molecules from the
brain and it regulates the influx of necessary nutrients, signaling molecules and immune cells into
the brain. The diverse processes governing the influx of necessary molecules for brain homeostasis
provide a variety of options that can be hijacked to improve the delivery of therapeutics and
diagnostics into the brain. These processes must be carefully engineered for safety and efficacy due
to the fragile nature of the brain [71-73].
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
340
Figure 2. Schematic summary of the possible routes across the blood brain barrier.
Reprinted by permission from Macmillan Publishers Ltd: Nature Reviews Neuroscience:
Abbott NJ, Ronnback L, Hansson E (2006) Astrocyte-endothelial interactions at the
blood-brain barrier. Nat Rev Neurosci 7: 41-53, copyright © 2006.
The essential function of the BBB is to maintain brain homeostasis. It can be characterized
through two routes: influx and efflux. The biological properties of the BBB that give it high
selectivity are as follows [73]:
a) The BBB has an extremely high resistivity as measured using transendothelial electrical
resistance (TEER). This is a measure of the resistance the BBB poses to the flow of charged ions. It
has a value between 1500–2000 Ω/m2 and is a result of the strong barrier posed by TJs between the
cells [73]. This resistance prevents access to most charged moieties through TJs so that only small,
neutral and water-soluble molecules can pass through these paracellular junctions.
b) The endothelial cells that make up the endothelium/capillary walls of the BBB is supported
by astrocytes and pericytes that provide polarization and communication with the endothelial cells to
adjust the local properties of the BBB as required, either up-regulating or down-regulating protein
and cell membrane composition [74].
c) Various transporter proteins are expressed at the endothelium including glucose carrier
(GLUT1) and amino acid carrier (LAT1) for transport of nutrients and small molecules needed for
brain metabolism. The transporter protein (cationic-organic transporter protein) is responsible for
shuttling small cationic species in and out and are implicated in the transport of current Alzheimer's
disease treatments memantine [75] and cholinesterase inhibitors [76] across the BBB.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
341
d) A family of efflux transporters are expressed at the endothelium such as P-glycoprotein
(P-gp), multidrug resistance protein (MRP) and receptor for advanced glycation end products
(RAGE), along with a class of lipoprotein receptors that are responsible for preventing entry into the
brain and actively transporting unwanted material out. RAGE and lipoprotein receptors appear to be
mainly responsible for transporting amyloid [77,78].
e) Receptor proteins for influx into the brain are also expressed on the blood side of the brain,
for instance: insulin receptors, transferrin receptors (TfR), lipoprotein receptors, diptheria toxin
receptors (DTR) and others. These transport larger molecules (mostly peptides, proteins and lipids)
into the brain.
f) Strict entry into the brain by immune cells is also permitted by the BBB, for instance
perivascular macrophages and monocytes [79].
3.2. Transport across the blood brain barrier with respect to AD pathophysiology
The aforementioned features combine to establish a strong physical, transport, metabolic and
immunologic barrier. To make use of any of these transport systems for increased uptake of drugs
across the BBB, it is important to consider the pathophysiology of the disease. There is no
one-size-fits-all drug delivery system for delivery of treatments to all the various neurodegenerative
and neurovascular disorders, cancers and infections associated with the brain. This is due to
differences in pathophysiology and BBB functionality caused by various diseases and conditions.
For example, transporter/receptor proteins can be expressed at different levels in a variety of
diseases and TJs in some disorders do not have as high a resistivity as in others. There are many
other factors that affect drug delivery in specific conditions. For a more general look at delivery
across the blood brain barrier, see Chen & Liu's paper from 2012 [71].
3.2.1. Paracellular and Transcellular routes
Only a small set of water soluble molecules and therapeutic drugs are able to pass through TJs
(shown in Figure 2a). Utilizing this form of transport into the brain for larger molecules relies on
generating transient reversible openings in the TJs between endothelial cells. This can be
accomplished through biological, chemical and physical means. Viruses are known to up-regulate
cytokines to disrupt the BBB and inflammatory stimuli such as histamine can increase blood-brain
permeability. Chemical stimuli, such as cyclodextrin can sequester cholesterol increasing BBB
permeation. Physically, it is possible to thermally excite the endothelium using low energy radiation
to increase permeation of the BBB (i.e. microwave and ultrasound waves) [71]. These methods are
non-specific and will increase the diffusion of most solutes from the blood into the brain.
This route into the brain would not be suitable for a sustained therapeutic treatment over an
extended period of time. Chronic conditions like AD would likely require continued treatment from
onset until the end of life. Opening the TJs compromises the integrity of the BBB which is already
been shown to be impaired [80-83]. The utilization of glucose, insulin receptor distribution and
efflux of Aβ are all known to be impaired in AD [84]. Considering this lack of amyloid efflux, it may
be beneficial in AD to re-establish a more effective, blood brain barrier rather than compromise
it [83-85]. In fact, many researchers consider neurovascular mechanisms of neurodegeneration in
Alzheimer's disease as highly relevant and even causative [84,85]. Ultimately, the cost-to-benefit
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
342
ratio is far too high for disrupting the BBB in AD.
Aside from attempting to pass drugs between cells through the TJs and AJs, molecules can also
be passed into the brain through the endothelial cells which make up the BBB. Passive transport
occurs for lipophilic molecules and important nutrients through carrier protein pathways. To
accommodate larger molecules the cell uses a vesicle mediated transport system called transcytosis.
It begins with endocytosis at the luminal (blood) side of the BBB. Here, receptor or electrostatic
interactions at the cell membrane trigger vesicle inclusion of the drug or molecule and is then shuttle
through the cytoplasm of the brain endothelial cell. Such inclusion of the drug molecule into the
vesicle protects it from endogenous enzymes. Following this, the molecule must then undergo
exocytosis at the aluminal (brain) side of the BBB. Drugs and molecules using transcellular routes
into the brain can do so via various routes, as summarized in Figure 2b‒e.
3.2.2. Lipophilic and transport protein pathways (influx/efflux)
Approximately 2% of drugs on the market are able to effectively cross the BBB [42]. Typically,
these molecules are small (less than 400 Da), lipophilic, neutral molecules capable of dissolving into
the cell membrane, like alcohols and steroidal hormones (see Figure 2b). This is a concentration
dependent route into the brain and by itself cannot support delivery of larger macromolecule
therapeutics such as peptides and MAbs for targeting amyloid.
As mentioned previously, endothelial cells express transporter molecules for moving glucose,
amino acids, nucleotides and other select small molecules into the brain for metabolic needs and
signaling pathways (Figure 2c). These transporter proteins are too small to deliver larger
macromolecules and nanoparticles into the brain, and their structure and functionality is not well
established. To effectively transport drugs using this technique, they must mimic the native nutrient's
molecular structure and be about the same size. These transporter proteins move essential nutrients
into the brain and trying to exploit them for drug delivery could have negative effects on brain
metabolism by competing with essential molecules [71].
Efflux pumps are a family of transporter proteins and receptors expressed in order to move
unwanted molecules out of the brain. They are expressed on both sides of the endothelium. They can
prevent passage of molecules on the blood side and actively transport them out from the brain side.
It is known that amyloid beta binds various efflux and influx pumps [77,78]. For the purpose of
delivering and keeping high concentrations of a drug to the brain, one can attempt to design the drug
as to not bind P-gp and other efflux pumps. If that is not possible it may be beneficial to utilize
efflux pump inhibitors. This will increase concentrations of the drug in the brain by limiting binding
events of the drug with the efflux pump. In an AD mouse model it was shown that P-gp deficiency
resulted in an increase in amyloid [77]. This suggests that inhibiting P-gp may increase amyloid
deposition. Generally, efflux inhibition will interfere with the natural protection offered by the efflux
pumps and may not be suitable for treating AD.
3.2.3. Receptor-mediated transcytosis
Receptor-mediated transcytosis (RMT) is a highly specific route across the BBB (Figure 2d).
Endothelial cells on the luminal side of the BBB express many cell surface proteins that bind many
different ligands including growth factors, hormones, enzymes and other substrates necessary for
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
343
brain metabolism. Recent advances in molecular biology have made it clear that attaching ligands
for receptors that are expressed at the BBB is a feasible way to have selective targeting to the brain.
With advances being made in proteomics and genetics it has been possible to more accurately
document the pathophysiology of the BBB in various disease states. This helps in narrowing down
suitable receptors for targeted delivery to the brain with optimal efficiency for a particular disease.
The understanding of expression of proteins at the BBB is not fully developed in the diseased brain,
but progress has been made, below are several examples of receptors related to AD pathology:
a) Transferrin Receptor (TfR). Transferrin binds free iron in the blood and the TfR is
responsible for transporting these loaded transferrin proteins across various tissue barriers in the
body, including the BBB. TfR density in the hippocampus of AD patients is decreased while it is
unchanged in cerebral micro-vessels [86], making it a suitable candidate for receptor mediated
delivery of drugs in AD. This receptor is a primary target used in RMT nanosystems in the literature.
Using transferrin itself as a ligand is not suitable since transferrin receptors are fully saturated by
native transferrin in circulation [87]. To overcome this, MAbs and other peptides directed against
distinct epitopes of the TfR, which do not compete with binding of native transferrin, have been
conjugated to drugs and delivery systems and used to target the brain [87-90]. Of these, the OX26
mouse MAb is the best studied. When conjugated to liposomes and other nanovehicles, it has been
shown to improve permeation across the BBB, both in vitro and in vivo [89]. TfR are heavily
expressed at the BBB but in other organs as well including liver, lung and kidney [88]. A rat MAb
directed against mouse TfR called R17-217 also undergoes RMT at the BBB (brain uptake at 1.7%
dose/g) with very low uptake in kidney or liver when compared to another MAb 8D3, which had a
higher brain uptake of 3.1% dose/g [89]. This suggests that the extracellular regions of TfR are
different in various cell types and that choosing the correct epitope may allow for targeting to the
brain with higher selectivity than other tissue types. This will reduce peripheral tissue uptake and
increase brain selective uptake.
b) Insulin receptors are important for maintaining glucose homoeostasis in the brain. As Aβ
binds readily to insulin receptors [91], glucose utilization is impaired due to the competition of
amyloid and insulin with the insulin receptor in AD patients [92]. This receptor may not be an ideal
candidate since Alzheimer's patients would not benefit from further glucose impairment.
c) Lipoprotein receptors are a broad class of receptor proteins that are responsible for
scavenging and signaling in the brain. It is known that down-regulation of low density lipoprotein
receptor related protein (LPR-1) decreased clearance of Aβ molecules from the brain. Competition
with an external ligand for drug delivery may impede the clearance of Aβ from the brain, which is
presumably detrimental in AD. On the flip side, another low density lipoprotein, the ApoE receptor
is responsible for shuttling cholesterol into the brain and thus has been suggested as a target for
improved BBB permeation [93,94].
d) Diphtheria toxin receptor (DTR) may be a good candidate for use in AD BBB delivery since
it has no ligands necessary for brain homoeostasis [95]. DTR is also known to be up-regulated in
conditions of inflammation in the brain which is associated with AD [95]. This may be a useful
target for both imaging and therapeutic delivery in AD. Due to the toxicity of the diphtheria toxin, it
is not a suitable ligand for drug delivery. A non-toxic mutant CRM197 has been explored for
targeting the DTR to improve drug delivery. It has been used as a component of vaccines to increase
immune response since the 1980s and therefore has a track record of being safe [71]. The transport
capacity using CRM197 was tested using a protein tracer across an in vitro model and in vivo using
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
344
guinea pigs. They found that transcytosis occurred at a slow rate after an initial delay. It is suggested
that antibody to CRM197 may be present in serum; if so it would neutralize its ability to bind to the
DTR [95]. That being said, it may be possible to target other epitopes of DTR using shorter
non-immunological peptide sequences.
e) Nicotinic acetylcholine receptors are the target of rabies virus glycoprotein (RVG), a 39
amino acid peptide. RVG binds the alpha-7 subunit, which is expressed in neurons and endothelial
cells; as such it has been widely explored as a brain targeting ligand especially in RNA interfering
strategies [96,97].
Receptor-mediated transcytosis has great potential for brain specific targeting of drug delivery
systems due to its specificity and increased BBB permeation. However, high affinity does not
guarantee highly efficient transcytosis. Yu et al. showed that a high affinity anti-TfR antibody had
lower levels of transcytosis than the lower affinity versions [66]. This implies that there is an ideal
affinity that will achieve a balance between selectivity and capacity. In a study by Couch et al.,
bispecific MAbs directed at TfRs and BACE1 with reduced affinity showed improved brain uptake
and safety while reducing peripheral exposure [98]. This same group was able to use imaging and
colocalization experiments to determine the fate of MAbs with various affinities, they found that
high affinity MAbs were trafficked to lysosomes for degradation and also significantly decreased
TfR activity reducing further uptake of new doses [99]. TfR/BACE1 bispecific antibodies showed
increased BBB penetration and reductions in brain amyloid compared to BACE1 MAb alone in
non-human primates; as well they showed that there is an ideal TfR affinity that is neither too high
nor too low [100]. Similarly, MAb directed against A have been conjugated with single Fab MAb
fragments that target TfRs and have demonstrated much improved BBB permeation. This
monovalent targeting of TfRs caused increases in brain deposition 55-fold while in contrast using
bivalent (two Fab MAb fragments), which increases affinity, resulted in lysosome trafficking and
reduced BBB permeation [101].
If a receptor is necessary for transporting essential nutrients or regulating brain homoeostasis, it
may not be a good target for RMT as competition in the compromised disease state may inflict more
harm than the drug will benefit. Therefore it is recommended that the ligand bind a different portion
of the receptor than the native substrate. A great challenge in determining the ideal receptor targeting
ligand is the lack of consistency in the system being tested and the model used for assessing BBB
permeation. In a step forward, five different RMT targeting ligands were tested in parallel using
identical liposomes in vivo. In comparing these five ligands only one MAb, R17-217 had sufficient
increases in BBB permeation at all of the time points while the other receptor ligands did not greatly
affect liposome delivery [102]. The results are summarized in Figure 3.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
345
Figure 3. Distribution of liposomes with various targeting ligands for RMT A) in the
body and B) in the brain at 12 hours. ID stands for injected dose. Reprinted from
Journal of Controlled Release, 150/1, van Rooy I, Mastrobattista E, Storm G, et al.
Comparison of five different targeting ligands to enhance accumulation of liposomes into
the brain, 30-36, Copyright 2011 with permission from Elsevier.
Nanoparticle delivery systems using receptor mediated transcytosis
In a study performed by Prades et al., they used a ligand for TfR to improve the delivery of
gold nanoparticles across the BBB. The gold nanoparticles (AuNP) were conjugated to a β-sheet
breaker peptide six amino acids long (CLPFFD) [90]. These AuNP-CLPFFD are able to recognize
amyloid and after irradiation with weak microwaves break up amyloid deposits as a form of
molecular surgery. These AuNP-peptide nanocomposites are unable to cross the BBB. To improve
permeation through the BBB they conjugated a peptide ligand for TfR. This short 12 amino acid
sequence (THRPPMWSPVWP) may be advantageous over MAb as it has high specificity and small
size (necessary to prevent steric hindrance of the small amyloid recognition peptide). The
nano-vehicles and controls are under 20 nm, which is favourable, unlike their highly negative
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
346
ζ-potential that contributes to reduced BBB permeation. To assess BBB permeation, a model
comprising bovine-brain endothelial cells (BBEC) on top of a filter, with rat astrocytes grown on the
bottom was used (Figure 4). This creates a supported and polarized model BBB with the top
compartment acting as the luminal (blood) side and the bottom acting as the aluminal (brain) side of
the BBB. The BBB permeation of the nanosystem can then be tested by introducing them in the top
compartment and measuring the amount in the bottom over time. The TEER of this model is about
an order of magnitude less than what the human BBB is expected to be in vivo, this is indicative of
adequate TJ formation for this type of in vitro assessment. By conjugating the nanocarrier with THR
(the ligand for TfR), an increase in BBB permeation was observed, by a factor of almost 1200 and a
factor of 12 in the presence fetal bovine serum. Passive diffusion is not possible as gold
nanoparticles were unable to permeate in a parallel artificial membrane permeability assay, PAMPA
for short.
Figure 4. A) In vitro model to test the efficacy of AuNP-THR-CLPFFD and control
nanocarrier's ability to permeate the BBB. B) PAMPA to test BBB permeation
through passive diffusion. Reprinted from Biomaterials, 33/29, Prades R, Guerrero S,
Araya E, et al.. Delivery of gold nanoparticles to the brain by conjugation with a peptide
that recognizes the transferrin receptor, 7194-7205, Copyright 2012 with permission
from Elsevier.
Lastly, in vivo assessments of BBB permeation on rats was performed. In this portion of the
study the rats were given a peritoneal injection of the THR-CLPFFD conjugated or control AuNPs.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
347
After, the animals were analyzed postmortem at 30 minute, 1 hour and 2 hour time intervals. The
gold content of the brain and peripheral organs was assessed using neutron activation. They found
high levels of gold in the liver and spleen, as to be expected (38 % and 18% of ID, respectively);
low levels of gold accumulation in the brain (0.07% of the ID) was sustained for at least an hour.
This was twice as high as the control AuNP accumulation after 1 hour. It was shown that an increase
in the uptake of nanoparticles can be achieved using TfR ligands, despite unfavorable electrostatic
surface potential. It's safe to suggest that a system with favorable energetics would have far
improved uptake at the BBB, compounding and complementing RMT.
In a recent study, Zhang et al. produced a dually functional nanosystem for delivery across the
blood brain barrier and targeting of Aβ [103]. They produced two peptide ligands, one for TfR and
one for Aβ, using phage display. These peptide ligands were then conjugated to the surface of a
polyethylene glycol-poly lactic acid polymer (PEG-PLA) nanoparticle. These nanoparticle systems
were approximately 100 nm in size, with Zeta (ζ) potential around −22 mV. MTT assays showed
negligible cytotoxicity while in vivo BBB permeation studies involving coumarin-6 dyes under
confocal microscope showed an increase in deposition in the brain [103]. In vivo studies focused on
blood brain barrier permeation for the purpose of evaluating depositions in the brain. Since no drug
or diagnostic payload was delivered, it was not necessary to monitor long term cognitive effects.
This study further highlights functionalized nanoparticles with peptides directed at TfR as an
effective strategy to promote increased BBB permeation in Aβ challenged brains. The authors
suggest that such a system may be useful to deliver a diagnostic or therapeutic. In a follow up study
this dually functional nanosystem was used to deliver a generic beta sheet breaker peptide, called
H102, to mice that had received intracranial injections of Aβ. A "sham" group received intracranial
injections of isotonic saline and was used as control [104]. The H102 bi-functional NPs exhibited
neuroprotective effects with two key biochemical indexes being restored to sham levels and
histology revealing reductions in pathology and cell loss in the hippocampus, while improved spatial
learning and memory was seen in the Morris water maze.
Based on earlier work done with lipid nanoparticles containing phosphatidic acid for targeting
amyloid [19], similar nanoliposomes were modified with lipoprotein receptor ligands (ApoE
peptides) to improve BBB permeation [94]. Later, this same group, again using phosphatidic acid
NPs to target amyloid, compared the effects of various linkages of MAbs directed at transferrin
receptors for improving BBB penetration. They showed that covalent linkage of anti-TfR MAb
R17-217, using Mal-PEG, to the liposome improves the ability of the liposome to breach the BBB as
compared to a biotin-streptavidin conjugation [105]. A different group used dually decorated
nanoliposomes with Aβ MAb and OX26 anti-TfR MAb using biotin streptavidin conjugation for
improved delivery across the blood brain barrier [106] (Figure 5 below). Subsequently, this group
tested anti-TfR MAb and an ApOE derived peptide as dual targeting ligands for improved absorption
in vitro, with a BBB model and in vivo animal studies. They found that both receptor ligands
improved uptake at the endothelium and had a compounding effect in vitro but not in vivo. These
conflicting reports were reconciled when serum was added to the cellular BBB model, at which
point they found ApoE was ineffective at improving uptake in the endothelial cells [93]. This may be
due to competitive processes between serum proteins and target receptor, or because serum proteins
directly quench ApoE peptides. These reports suggest that covalent conjugation of the targeting
ligands are superior to biotin-streptavidin systems and that transferrin receptor ligands may be an
ideal choice of targeting ligand over ApoE peptides.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
348
Figure 5. Dually decorated nanoliposomes targeting both TfR and Aβ with MAb.
Reprinted from European Journal of Pharmaceutics and Biopharmaceutics, 81/1,
Markoutsa E, Papadia K, Clemente C, et al. Anti-Abeta-MAb and dually decorated
nanoliposomes: effect of Abeta1-42 peptides on interaction with hCMEC/D3 cells, 49-56,
Copyright 2012 with permission from Elsevier.
3.2.4. Absorptive-mediate transcytosis
Native positively charged proteins and macromolecules such as human serum albumin are
capable of transcytosis across the BBB through electrostatic interactions with the negatively charged
micro domains in the endothelial cell membrane; this process is called absorptive -mediated
transcytosis or AMT (Figure 2e). By enhancing the surface of nanoparticle drug delivery systems
with positively charged moieties, AMT has the potential to greatly improve the permeation of many
drugs at the BBB. AMT begins with a non-specific electrostatic interaction with the endothelial cell
membrane and as such has a high capacity for improving BBB permeation [107]. As a result of the
low specificity of AMT these drug delivery system will have increased uptake in other tissues in the
body, such as the liver and kidney. The increase in permeation into other organ systems leads to less
drug delivered to the brain. This challenge is not insurmountable as it is known that polyethylene
glycol (PEG) coated nanoparticles have decreased liver uptake and increased blood circulation
time [18]. Moreover, by combining AMT favorable surface enhancement with RMT, it may be
possible to have efficient, high capacity, and high specificity drug delivery across the BBB for the
treatment of various disorders including AD.
Several studies have shown that molecules with positive moieties and positive ζ potentials can
increase permeation through the blood brain barrier [107]. Protamine [108], cationic proteins [109],
and a class of cell-penetrating peptides (CPPs) [110] have been used to improve AMT. CPPs are an
especially versatile class of short (less than 30 amino acids long) peptides capable of improving
penetration into cells. The exact mechanisms governing CPP are currently in debate and, in fact,
many mechanisms may be acting in parallel; evidence suggests cationic CPPs may involve
endocytosis pathways as in AMT [110]. A CPP derived from the HIV trans-activator of transcription,
TAT for short, was retro inversed and conjugated to the previously mentioned RI-OR2
anti-aggregation inhibitor with remarkable results in an in vivo transgenic mouse model [65]. The
TAT/RI-OR2 peptide was able to cross the blood brain barrier and bind plaques, after 21 days of
intraperitoneal injections dual transgenic mice had 25% reduction in cerebral cortex Aβ, 32%
reduction in plaque count, 44% reduction in activated microglial cells, 25% reduction in oxidative
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
349
damage and 210% increased neuron count in the dentate gyrus. These results indicate that oxidative
damage, reduced neurogenesis and inflammation are results of Aβ aggregation and can be reduced
with the use of peptide inhibitors provided they can be delivered to the brain.
Aside from CPPs, nanostructures which contain a positive charge hold great promise as either
modifications to surface enhanced drug delivery systems or as the core nanomaterial themselves.
Cationic polymers, typically featuring amine groups, have been shown to have to increase
permeation at the BBB. This is due to the protonation of these groups in neutral to acidic conditions
such as physiological pH [71,111]. Polycationic polymers are known to be cytotoxic, are not
typically suitable for clinical applications, and their mechanism is not well understood, though both
necrotic and apoptotic factors are involved [112]. One natural cationic polymer stands out, several
studies have shown that chitosan conjugated nanoparticles can increase uptake at the BBB and do so
safely [111,113]. Chitosan is a natural polymeric molecule comprised of randomly distributed
polysaccharides containing amine groups and has been of great interest as a nanoparticle drug carrier
in many diseases including AD [114]. These amine groups become positively protonated in
physiological pH this generates a favorable surface potential for endosome forming cell membrane
interactions. Many groups have utilized chitosan as a safe and effective polymeric delivery
vehicle [111,113,115,116].
Nanosystems utilizing absorptive-mediated transcytosis to cross the BBB
One model nano-vehicle consists of a polylactic-co-glycolic acid (PLGA) polymer nanocore,
surface enhanced with chitosan for improved electrostatic surface potential to facilitate AMT and
anti-amyloid IgG4.1 antibody for targeting of amyloid deposits [111]. PLGA is a biodegradable,
biocompatible, non-toxic, non-immunogenic copolymer of lactic and glycolic acid. The ratio of
lactic to glycolic acid in this co-polymer change its rate of degradation and in a drug delivery system
will allow for a tunable drug release profile. A higher ratio of glycolide to lactide lowers the
degradation time of PLGA, with the exception of a 50:50 ratio which leads to fastest degradation [111].
The surface enhancing molecule chitosan is a naturally occurring polysaccharide; it is made of
randomly distributed β-linked ᴅ-glucosamine (deacetylated unit) and N-acetyl-ᴅ-glucosamine
(acetylated unit). Since chitosan contains amine groups, it becomes protonated in acidic to neutral
conditions. Its ζ potential is around 20 mV at a pH of 5 and around 10 mV at a pH of 7
(physiological conditions). This positive ζ potential contributes to interactions at the endothelial cell
membrane and also improves its colloidal stability. Chitosan's properties can be modified in many
ways and also satisfies the conditions required by nanoparticle drug delivery systems with regards to
biocompatibility, biodegradation, toxicity and immunogenicity. Chitosan was also shown to be an
effective cryo-protectant, which is important for long-term stability during storage of drugs and
therapeutics.
Using a polarized Madin-Darby canine kidney (MDCK) cell monolayer model to assess BBB,
an increase in the uptake of immuno-nanovehicles conjugated with chitosan verses the control was
found [111]. This shows that chitosan can induce AMT at the BBB. Moreover, cells challenged with
Aβ had a very large increase in the uptake of immuno-nanovehicles, whether this is due to targeting
of Aβ or because Aβ reduces the integrity of the BBB remains to be seen. It is likely that both may
factor into the total BBB permeation. These results were visually demonstrated using confocal laser
microscopy shown in Figure 6. The utility of using chitosan nanostructures for improved AMT
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
350
across the BBB, for targeting Aβ, has been shown. However, verification of its ability to sufficiently
cross the BBB in vivo is necessary. Since this nanovehicle does not have selectivity for brain and its
surface potential will likely result in low blood circulation time and therefore a reduced uptake at the
BBB due to clearance of the nanovehicle in the liver, kidneys and lungs. Possible improvements
would include adding targeting ligands for the brain and PEG conjugation to reduce liver uptake.
This PEG conjugation, however, is associated with a lowering of ζ potential. More work needs to be
done to verify whether or not this system would achieve sufficient concentrations in vivo, and
characterize the deposition in other body tissues.
(green dye). B) Control
Figure 6. Laser confocal microscopy images of MDCK cell monolayer under various
immuno-nanovehicles (IgG4.1 present)
test conditions. A) Uptake of CPLGA
encapsulated with 6-coumarin
immune-nanovehicles
(PLGA-IgG4.1, no chitosan present), notice the lack of 6-coumarin, implies low uptake.
C) 6-coumarin chitosan conjugated PLGA nanoparticle (no IgG4.1 present). D)
Immuno-nanovehicles
in MDCK monolayer, note moderate absorption. E)
Immuno-nanovehicles in MDCK monolayer challenged with Aβ, note large amounts of
endocytosis and transcytosis (columns of bright green). Reprinted from Nanomedicine:
Nanotechnology, Biology and Medicine, 8/2, Jaruszewski KM, Ramakrishnan S, Poduslo
JF, et al. Chitosan enhances the stability and targeting of immuno-nanovehicles to
cerebro-vascular deposits of Alzheimer's disease amyloid protein, 250-260, Copyright
2012 with permission from Elsevier.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
351
Polymer-based nanoparticle drug delivery systems are readily modified. Chitosan PEG
copolymers represent a balance between favorable blood circulation time and ζ potential, especially
for the delivery of gene therapies and negatively charged peptides. Moreover, there are safety
concerns associated with cationic polymer nanoparticles as they may cause disruption of the cell
membrane. The use of PEG to quench some of this positive charge may improve safety concerns.
Malhotra et al. were able to conjugate up to 7.5% of the amine groups of chitosan with PEG,
forming nanoparticles less than 100 nm [113]. Afterwards, they used these PEG molecules as linkers
for a cell-penetrating peptide to increase AMT. They showed this significantly improved uptake of
siRNAs delivered to neuroblastomas in vitro.
4. Future directions
The delivery of therapeutics across the BBB for the treatment of Alzheimer's disease can be
improved using nanoparticles. The high capacity utility of AMT and the specificity of RMT are both
highly desirable to efficiently deliver compounds to the brain. The improved delivery of drugs across
the BBB has been shown through the use of many different systems and tested using a wide array of
methodologies, in vitro and in vivo. All of these studies show promise, however the lack of
consistent approaches has made it difficult to track down any one best method for a given disorder.
That being said, it is still possible to use these preliminary studies to think critically about novel
systems that will have improved efficacy over previous ones. These nanoparticle drug delivery
systems use a basic nanocore structure (amphiphilic, polymeric, or metallic), a targeting ligand for
Aβ and some other surface enhancement for BBB permeation, either AMT or RMT. Utilizing both
AMT and RMT to improve the surface characteristics of the nanoparticle may improve the
efficiency of BBB permeation for delivering therapeutics.
Liposomes can go through bi-ligand modification for delivery to neurons that utilizes both TfR
targeting with native transferrin (for RMT) to improve specificity and a poly-ʟ-arginine CPP to
improve capacity through AMT [116]. In this particular example, the β-galactosidase reporter gene is
delivered in vivo to adult rats. The liposomes were conjugated with poly-ʟ-arginine, which is
protonated in most acidic, neutral, and basic conditions. This resulted in a final ζ potential around 10
mV for bi-ligand nanovehicle. The TfR ligand used was found to reduce the zeta-potential by
approximately 10 mV. They were able to achieve a deposition of 4% of the injected dose per gram of
tissue at a 24 hour time point, which was twice the injected dose (2% ID/mg tissue) as seen in the
transferrin modified liposome alone. Less than 1% was seen in the control liposome. A detailed
analysis of the deposition of bi-ligand liposomes in the various organs was performed, with
considerable uptake in the liver, spleen, kidney, lungs and heart. This is not surprising as transferrin
receptors are expressed in all these tissue types. This system may be improved by utilizing a peptide
or MAb ligand more specific to brain TfR which does not complete with native transferrin
homeostasis, such as R17-217 MAb.
Another system that combines RMT and AMT pathways for improved uptake at the BBB is an
RVG-peptide-linked trimethylated chitosan for delivery of siRNA against BACE1 transcripts to the
brain [115]. RVG is known to bind nicotinic acetylcholine receptor subunit alpha-7 expressed at the
luminal side of the endothelium as well as on neurons. This improves uptake at the BBB and at the
neuronal cell membrane. The nanocore was built out of trimethylated chitosan combined to form a
copolymer with PEG, as similar to what was seen previously. The final siRNA delivery vehicle had a
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
352
ζ potential that decreased with the ratio of positively charged amino groups on chitosan to negatively
charged siRNA. At a ratio of 96:1 the zeta potential was 9 ± 2.5 mV for good AMT capacity. This
system benefits from the serum stability, and thus longevity, afforded by PEG and has the targeting
capability of RVG. In vitro gene silencing showed a 50% reduction in BACE1 levels while in vivo
imaging showed significant deposition in the brain compared to background levels found in the
control, simultaneously deposition in the kidney and liver decreased slightly. There was no
deposition found in the heart or spleen and hepatotoxicity was not observed. This study shows
promise for the development of more efficient drug delivery vehicles. Long-term studies involving
behavior and effect on cognition in AD mouse models are necessary to verify the safety and efficacy
of this BACE1 delivery system.
5. Conclusion
Over 100 years have passed since the first characterization of AD and although significant
progress in the molecular mechanisms has been seen, there is still much that needs to be clarified. Of
particular importance is the verification of the amyloid cascade hypothesis. It has already been
determined that Aβ directed therapeutics are not sufficient to treat mild to moderate AD and a shift in
paradigms to preventative clinical trials has begun in the last couple of years. It is important to note
that many current strategies being explored in preventive clinical trials were not designed to
intervene in a preventative manner and that other therapeutics may offer significant advantages.
Moreover, a null result may still be seen if efficient delivery of therapeutics to the brain cannot be
achieved. Regardless of the success or failure of the amyloid cacscade hypothesis, better drug
delivery systems for improving delivery of pharmaceuticals to the brain are still a priority and
lessons learned from targeting Aβ can be applied to the delivery of other drugs for treating AD
should they prove more relevant in the future.
Acknowledgments
The authors thank Dr. E. Drolle and Dr. R. Henderson for helpful discussion and Dr. F. Hane for
critical reading of the final manuscript. This work has been funded by Natural Science and
Engineering Council of Canada (NSERC) and University of Waterloo Chronic Disease Prevention
Initiative (CDPI) grants to ZL.
Conflict of interest
Authors declare no conflict of interest.
References
1. Prince M, Jackson J (2009) Alzheimer's Disease International World Alzheimer Report 2009.
London. 1-96 p.
2. Takeda A, Loveman E, Clegg A, et al. (2006) A systematic review of the clinical effectiveness of
donepezil, rivastigmine and galantamine on cognition, quality of life and adverse events in
Alzheimer's disease. Int J Geriatr Psychiatry 21: 17-28.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
353
3. Ong KT, Villemagne VL, Bahar-Fuchs A, et al. (2015) Abeta imaging with 18F-florbetaben in
prodromal Alzheimer's disease: a prospective outcome study. J Neurol Neurosurg Psychiatry 86:
431-436.
4. Querfurth HW, LaFerla FM (2010) Alzheimer's disease. N Engl J Med 362: 329-344.
5. Sperling RA, Aisen PS, Beckett LA, et al. (2011) Toward defining the preclinical stages of
Alzheimer's disease: recommendations from the National Institute on Aging-Alzheimer's
Association workgroups on diagnostic guidelines for Alzheimer's disease. Alzheimers Dement 7:
280-292.
6. Petkova AT, Ishii Y, Balbach JJ, et al. (2002) A structural model for Alzheimer's beta -amyloid
fibrils based on experimental constraints from solid state NMR. Proc Natl Acad Sci U S A 99:
16742-16747.
7. Matsuzaki K (2014) How do membranes initiate Alzheimer's Disease? Formation of toxic
amyloid fibrils by the amyloid beta-protein on ganglioside clusters. Acc Chem Res 47: 2397-2404.
8. Cecchi C, Stefani M (2013) The amyloid-cell membrane system. The interplay between the
biophysical features of oligomers/fibrils and cell membrane defines amyloid toxicity. Biophys
Chem 182: 30-43.
9. Drolle E, Gaikwad RM, Leonenko Z (2012) Nanoscale electrostatic domains in cholesterol-laden
lipid membranes create a target for amyloid binding. Biophys J 103: L27-29.
10. Drolle E, Hane F, Lee B, et al. (2014) Atomic force microscopy to study molecular mechanisms
of amyloid fibril formation and toxicity in Alzheimer's disease. Drug Metab Rev 46: 207-223.
11. Hane F, Drolle E, Gaikwad R, et al. (2011) Amyloid-beta aggregation on model lipid membranes:
an atomic force microscopy study. J Alzheimers Dis 26: 485-494.
12. Lal R, Lin H, Quist AP (2007) Amyloid beta ion channel: 3D structure and relevance to amyloid
channel paradigm. Biochim Biophys Acta 1768: 1966-1975.
13. Arispe N, Pollard HB, Rojas E (1993) Giant multilevel cation channels formed by Alzheimer
disease amyloid beta-protein [A beta P-(1-40)] in bilayer membranes. Proc Natl Acad Sci U S A
90: 10573-10577.
14. Demuro A, Smith M, Parker I (2011) Single-channel Ca(2+) imaging implicates Abeta1-42
amyloid pores in Alzheimer's disease pathology. J Cell Biol 195: 515-524.
15. Plant LD, Boyle JP, Smith IF, et al. (2003) The production of amyloid beta peptide is a critical
requirement for the viability of central neurons. J Neurosci 23: 5531-5535.
16. Giuffrida ML, Caraci F, Pignataro B, et al. (2009) Beta-amyloid monomers are neuroprotective. J
Neurosci 29: 10582-10587.
17. Hardy J, Selkoe DJ (2002) The Amyloid Hypothesis of Alzheimer's Disease: Progress and
Problems on the Road to Therapeutics. Science 297: 353-356.
18. Brambilla D, Verpillot R, Le Droumaguet B, et al. (2012) PEGylated nanoparticles bind to and
alter amyloid-beta peptide conformation: toward engineering of functional nanomedicines for
Alzheimer's disease. ACS Nano 6: 5897-5908.
19. Gobbi M, Re F, Canovi M, et al. (2010) Lipid-based nanoparticles with high binding affinity for
amyloid-beta1-42 peptide. Biomaterials 31: 6519-6529.
20. Cheng KK, Yeung CF, Ho SW, et al. (2013) Highly stabilized curcumin nanoparticles tested in an
in vitro blood-brain barrier model and in Alzheimer's disease Tg2576 mice. Aaps j 15: 324-336.
21. Cheng X, van Breemen RB (2005) Mass Spectrometry-Based Screening for Inhibitors of
β-Amyloid Protein Aggregation. Anal Chem 77: 7012-7015.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
354
22. He H, Dong W, Huang F (2010) Anti-Amyloidogenic and Anti-Apoptotic Role of Melatonin in
Alzheimer Disease. Curr Neuropharmacol 8: 211-217.
23. Lin L, Huang Q-X, Yang S-S, et al. (2013) Melatonin in Alzheimer's Disease. Int J Mol Sci 14:
14575-14593.
24. Rosales-Corral S, Acuna-Castroviejo D, Tan DX, et al. (2012) Accumulation of Exogenous
Amyloid-Beta Peptide in Hippocampal Mitochondria Causes Their Dysfunction: A Protective
Role for Melatonin. Oxid Med Cell Longev 2012: 843649.
25. Adolfsson O, Pihlgren M, Toni N, et al. (2012) An effector-reduced anti-beta-amyloid (Abeta)
antibody with unique abeta binding properties promotes neuroprotection and glial engulfment of
Abeta. J Neurosci 32: 9677-9689.
26. DeMattos RB, Bales KR, Cummins DJ, et al. (2001) Peripheral anti-A beta antibody alters CNS
and plasma A beta clearance and decreases brain A beta burden in a mouse model of Alzheimer's
disease. Proc Natl Acad Sci U S A 98: 8850-8855.
27. Legleiter J, Czilli DL, Gitter B, et al. (2004) Effect of different anti-Abeta antibodies on Abeta
fibrillogenesis as assessed by atomic force microscopy. J Mol Biol 335: 997-1006.
28. Lemere CA (2013) Immunotherapy for Alzheimer's disease: hoops and hurdles. Mol
Neurodegener 8: 36.
29. Robert R, Wark KL (2012) Engineered antibody approaches for Alzheimer's disease
immunotherapy. Arch Biochem Biophys 526: 132-138.
30. Wilcock DM, Rojiani A, Rosenthal A, et al. (2004) Passive immunotherapy against Abeta in aged
APP-transgenic mice reverses cognitive deficits and depletes parenchymal amyloid deposits in
spite of increased vascular amyloid and microhemorrhage. J Neuroinflammation 1: 24.
31. Findeis MA, Musso GM, Arico-Muendel CC, et al. (1999) Modified-peptide inhibitors of
amyloid beta-peptide polymerization. Biochemistry 38: 6791-6800.
32. Ghanta J, Shen CL, Kiessling LL, et al. (1996) A strategy for designing inhibitors of beta-amyloid
toxicity. J Biol Chem 271: 29525-29528.
33. Hane FT, Lee BY, Petoyan A, et al. (2014) Testing synthetic amyloid-beta aggregation inhibitor
using single molecule atomic force spectroscopy. Biosens Bioelectron 54: 492-498.
34. Pallitto MM, Ghanta J, Heinzelman P, et al. (1999) Recognition sequence design for peptidyl
modulators of beta-amyloid aggregation and toxicity. Biochemistry 38: 3570-3578.
35. Tjernberg LO, Naslund J, Lindqvist F, et al. (1996) Arrest of beta-amyloid fibril formation by a
pentapeptide ligand. J Biol Chem 271: 8545-8548.
36. Morgan D, Diamond DM, Gottschall PE, et al. (2000) A beta peptide vaccination prevents
memory loss in an animal model of Alzheimer's disease. Nature 408: 982-985.
37. Orgogozo JM, Gilman S, Dartigues JF, et al. (2003) Subacute meningoencephalitis in a subset of
patients with AD after Abeta42 immunization. Neurology 61: 46-54.
38. Wildsmith KR, Holley M, Savage JC, et al. (2013) Evidence for impaired amyloid beta clearance
in Alzheimer's disease. Alzheimers Res Ther 5: 33.
39. Evin G, Hince C (2013) BACE1 as a therapeutic target in Alzheimer's disease: rationale and
current status. Drugs Aging 30: 755-764.
40. Wolfe MS (2012) gamma-Secretase inhibitors and modulators for Alzheimer's disease. J
Neurochem 120 Suppl 1: 89-98.
41. Atwal JK, Chen Y, Chiu C, et al. (2011) A therapeutic antibody targeting BACE1 inhibits
amyloid-beta production in vivo. Sci Transl Med 3: 84ra43.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
355
42. Kao SC, Krichevsky AM, Kosik KS, et al. (2004) BACE1 suppression by RNA interference in
primary cortical neurons. J Biol Chem 279: 1942-1949.
43. Singer O, Marr RA, Rockenstein E, et al. (2005) Targeting BACE1 with siRNAs ameliorates
Alzheimer disease neuropathology in a transgenic model. Nat Neurosci 8: 1343-1349.
44. Vassar R (2014) BACE1 inhibitor drugs in clinical trials for Alzheimer's disease. Alzheimers Res
Ther 6: 89.
45. Luo Y, Bolon B, Kahn S, et al. (2001) Mice deficient in BACE1, the Alzheimer's beta-secretase,
have normal phenotype and abolished beta-amyloid generation. Nat Neurosci 4: 231-232.
46. Jacobsen H, Ozmen L, Caruso A, et al. (2014) Combined treatment with a BACE inhibitor and
anti-Abeta antibody gantenerumab enhances amyloid reduction in APPLondon mice. J Neurosci
34: 11621-11630.
47. Morgan D (2005) Mechanisms of A beta plaque clearance following passive A beta immunization.
Neurodegener Dis 2: 261-266.
48. Banks WA, Terrell B, Farr SA, et al. (2002) Passage of amyloid beta protein antibody across the
blood-brain barrier in a mouse model of Alzheimer's disease. Peptides 23: 2223-2226.
49. Banks WA (2012) Drug delivery to the brain in Alzheimer's disease: consideration of the
blood-brain barrier. Adv Drug Deliv Rev 64: 629-639.
50. Pfeifer LA, White LR, Ross GW, et al. (2002) Cerebral amyloid angiopathy and cognitive
function: the HAAS autopsy study. Neurology 58: 1629-1634.
51. Racke MM, Boone LI, Hepburn DL, et al. (2005) Exacerbation of cerebral amyloid
angiopathy-associated microhemorrhage in amyloid precursor protein transgenic mice by
immunotherapy is dependent on antibody recognition of deposited forms of amyloid beta. J
Neurosci 25: 629-636.
52. Wilcock DM, Colton CA (2008) Anti-amyloid-beta immunotherapy in Alzheimer's disease:
relevance of transgenic mouse studies to clinical trials. J Alzheimers Dis 15: 555-569.
53. Doody RS, Thomas RG, Farlow M, et al. (2014) Phase 3 trials of solanezumab for
mild-to-moderate Alzheimer's disease. N Engl J Med 370: 311-321.
54. Panza F, Solfrizzi V, Imbimbo BP, et al. (2014) Amyloid-based immunotherapy for Alzheimer's
disease in the time of prevention trials: the way forward. Expert Rev Clin Immunol 10: 405-419.
55. Salloway S, Sperling R, Fox NC, et al. (2014) Two phase 3 trials of bapineuzumab in
mild-to-moderate Alzheimer's disease. N Engl J Med 370: 322-333.
56. Tayeb HO, Murray ED, Price BH, et al. (2013) Bapineuzumab and solanezumab for Alzheimer's
disease: is the 'amyloid cascade hypothesis' still alive? Expert Opin Biol Ther 13: 1075-1084.
57. Garber K (2012) Genentech's Alzheimer's antibody trial to study disease prevention. Nat
Biotechnol 30: 731-732.
58. Panza F, Solfrizzi V, Imbimbo BP, et al. (2014) Efficacy and safety studies of gantenerumab in
patients with Alzheimer's disease. Expert Rev Neurother 14: 973-986.
59. Kumar J, Sim V (2014) D-amino acid-based peptide inhibitors as early or preventative therapy in
Alzheimer disease. Prion 8: 119-124.
60. Porat Y, Mazor Y, Efrat S, et al. (2004) Inhibition of islet amyloid polypeptide fibril formation: a
potential role for heteroaromatic interactions. Biochemistry 43: 14454-14462.
61. Austen BM, Paleologou KE, Ali SA, et al. (2008) Designing peptide inhibitors for
oligomerization and toxicity of Alzheimer's beta-amyloid peptide. Biochemistry 47: 1984-1992.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
356
62. Taylor M, Moore S, Mayes J, et al. (2010) Development of a proteolytically stable retro-inverso
peptide inhibitor of beta-amyloid oligomerization as a potential novel treatment for Alzheimer's
disease. Biochemistry 49: 3261-3272.
63. Roy S (2010) Designing Novel Peptidic Inhibitors of Beta Amyloid Oligomerization Calgary,
Alberta: University of Calgary. 164 p.
64. Gordon DJ, Tappe R, Meredith SC (2002) Design and characterization of a membrane permeable
N-methyl amino acid-containing peptide that inhibits Abeta1-40 fibrillogenesis. J Pept Res 60:
37-55.
65. Parthsarathy V, McClean PL, Holscher C, et al. (2013) A novel retro-inverso peptide inhibitor
reduces amyloid deposition, oxidation and inflammation and stimulates neurogenesis in the
APPswe/PS1DeltaE9 mouse model of Alzheimer's disease. PLoS One 8: e54769.
66. Yu YJ, Zhang Y, Kenrick M, et al. (2011) Boosting brain uptake of a therapeutic antibody by
reducing its affinity for a transcytosis target. Sci Transl Med 3: 84ra44.
67. Xiong N, Dong X-Y, Zheng J, et al. (2015) Design of LVFFARK and LVFFARK-functionalized
nanoparticles for inhibiting amyloid β-protein fibrillation and cytotoxicity. ACS Appl Mater
Interfaces 7: 5650-5662.
68. Taylor M, Moore S, Mourtas S, et al. (2011) Effect of curcumin-associated and lipid
ligand-functionalized nanoliposomes on aggregation of the Alzheimer's Abeta peptide.
Nanomedicine 7: 541-550.
69. Armstrong RA (2014) A critical analysis of the 'amyloid cascade hypothesis'. Folia Neuropathol
52: 211-225.
70. Castello MA, Soriano S (2014) On the origin of Alzheimer's disease. Trials and tribulations of the
amyloid hypothesis. Ageing Res Rev 13: 10-12.
71. Chen Y, Liu L (2012) Modern methods for delivery of drugs across the blood-brain barrier. Adv
Drug Deliv Rev 64: 640-665.
72. Pardridge WM (2003) Blood-brain barrier drug targeting: the future of brain drug development.
Mol Interv 3: 90-105, 151.
73. Abbott NJ, Patabendige AA, Dolman DE, et al. (2010) Structure and function of the blood-brain
barrier. Neurobiol Dis 37: 13-25.
74. Abbott NJ, Ronnback L, Hansson E (2006) Astrocyte-endothelial interactions at the blood-brain
barrier. Nat Rev Neurosci 7: 41-53.
75. Mehta DC, Short JL, Nicolazzo JA (2013) Memantine transport across the mouse blood-brain
barrier is mediated by a cationic influx H+ antiporter. Mol Pharm 10: 4491-4498.
76. Kim MH, Maeng HJ, Yu KH, et al. (2010) Evidence of carrier-mediated transport in the
penetration of donepezil into the rat brain. J Pharm Sci 99: 1548-1566.
77. Cirrito JR, Deane R, Fagan AM, et al. (2005) P-glycoprotein deficiency at the blood-brain barrier
increases amyloid-beta deposition in an Alzheimer disease mouse model. J Clin Invest 115:
3285-3290.
78. Deane R, Wu Z, Zlokovic BV (2004) RAGE (yin) versus LRP (yang) balance regulates alzheimer
amyloid beta-peptide clearance through transport across the blood-brain barrier. Stroke 35:
2628-2631.
79. Guillemin GJ, Brew BJ (2004) Microglia, macrophages, perivascular macrophages, and pericytes:
a review of function and identification. J Leukoc Biol 75: 388-397.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
357
80. Bowman GL, Kaye JA, Moore M, et al. (2007) Blood-brain barrier impairment in Alzheimer
disease: stability and functional significance. Neurology 68: 1809-1814.
81. Claudio L (1996) Ultrastructural features of the blood-brain barrier in biopsy tissue from
Alzheimer's disease patients. Acta Neuropathol 91: 6-14.
82. Kalaria RN (1992) The blood-brain barrier and cerebral microcirculation in Alzheimer disease.
Cerebrovasc Brain Metab Rev 4: 226-260.
83. Marco S, Skaper SD (2006) Amyloid beta-peptide1-42 alters tight junction protein distribution
and expression in brain microvessel endothelial cells. Neurosci Lett 401: 219-224.
84. Zlokovic BV (2005) Neurovascular mechanisms of Alzheimer's neurodegeneration. Trends
Neurosci 28: 202-208.
85. Wisniewski HM, Vorbrodt AW, Wegiel J (1997) Amyloid angiopathy and blood-brain barrier
changes in Alzheimer's disease. Ann N Y Acad Sci 826: 161-172.
86. Kalaria RN, Sromek SM, Grahovac I, et al. (1992) Transferrin receptors of rat and human brain
and cerebral microvessels and their status in Alzheimer's disease. Brain Res 585: 87-93.
87. Jones AR, Shusta EV (2007) Blood-brain barrier transport of therapeutics via receptor-mediation.
Pharm Res 24: 1759-1771.
88. Gatter KC, Brown G, Trowbridge IS, et al. (1983) Transferrin receptors in human tissues: their
distribution and possible clinical relevance. J Clin Pathol 36: 539-545.
89. Lee HJ, Engelhardt B, Lesley J, et al. (2000) Targeting rat anti-mouse transferrin receptor
monoclonal antibodies through blood-brain barrier in mouse. J Pharmacol Exp Ther 292:
1048-1052.
90. Prades R, Guerrero S, Araya E, et al. (2012) Delivery of gold nanoparticles to the brain by
conjugation with a peptide that recognizes the transferrin receptor. Biomaterials 33: 7194-7205.
91. Zhao WQ, De Felice FG, Fernandez S, et al. (2008) Amyloid beta oligomers induce impairment
of neuronal insulin receptors. Faseb J 22: 246-260.
92. Xie L, Helmerhorst E, Taddei K, et al. (2002) Alzheimer's beta-amyloid peptides compete for
insulin binding to the insulin receptor. J Neurosci 22: Rc221.
93. Markoutsa E, Papadia K, Giannou AD, et al. (2014) Mono and dually decorated nanoliposomes
for brain targeting, in vitro and in vivo studies. Pharm Res 31: 1275-1289.
94. Re F, Cambianica I, Sesana S, et al. (2010) Functionalization with ApoE-derived peptides
enhances the interaction with brain capillary endothelial cells of nanoliposomes binding
amyloid-beta peptide. J Biotechnol 156: 341-346.
95. Gaillard PJ, Visser CC, de Boer AG (2005) Targeted delivery across the blood-brain barrier.
Expert Opin Drug Deliv 2: 299-309.
96. Lentz TL (1990) Rabies virus binding to an acetylcholine receptor alpha-subunit peptide. J Mol
Recognit 3: 82-88.
97. Liu Y, Huang R, Han L, et al. (2009) Brain-targeting gene delivery and cellular internalization
mechanisms for modified rabies virus glycoprotein RVG29 nanoparticles. Biomaterials 30:
4195-4202.
98. Couch JA, Yu YJ, Zhang Y, et al. (2013) Addressing safety liabilities of TfR bispecific antibodies
that cross the blood-brain barrier. Sci Transl Med 5: 183ra157, 181-112.
99. Bien-Ly N, Yu YJ, Bumbaca D, et al. (2014) Transferrin receptor (TfR) trafficking determines
brain uptake of TfR antibody affinity variants. J Exp Med 211: 233-244.
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
358
100. Yu YJ, Atwal JK, Zhang Y, et al. (2014) Therapeutic bispecific antibodies cross the blood-brain
barrier in nonhuman primates. Sci Transl Med 6: 261ra154.
101. Niewoehner J, Bohrmann B, Collin L, et al. (2014) Increased brain penetration and potency of a
therapeutic antibody using a monovalent molecular shuttle. Neuron 81: 49-60.
102. van Rooy I, Mastrobattista E, Storm G, et al. (2011) Comparison of five different targeting
ligands to enhance accumulation of liposomes into the brain. J Control Release 150: 30-36.
103. Zhang C, Wan X, Zheng X, et al. (2014) Dual-functional nanoparticles targeting amyloid plaques
in the brains of Alzheimer's disease mice. Biomaterials 35: 456-465.
104. Zhang C, Zheng X, Wan X, et al. (2014) The potential use of H102 peptide-loaded dual-functional
nanoparticles in the treatment of Alzheimer's disease. J Control Release 192: 317-324.
105. Salvati E, Re F, Sesana S, et al. (2013) Liposomes functionalized to overcome the blood-brain
barrier and to target amyloid-beta peptide: the chemical design affects the permeability across an
in vitro model. Int J Nanomedicine 8: 1749-1758.
106. Markoutsa E, Papadia K, Clemente C, et al. (2012) Anti-Abeta-MAb and dually decorated
nanoliposomes: effect of Abeta1-42 peptides on interaction with hCMEC/D3 cells. Eur J Pharm
Biopharm 81: 49-56.
107. Herve F, Ghinea N, Scherrmann JM (2008) CNS delivery via adsorptive transcytosis. AAPS J 10:
455-472.
108. Pardridge WM, Buciak JL, Kang YS, et al. (1993) Protamine-mediated transport of albumin into
brain and other organs of the rat. Binding and endocytosis of protamine-albumin complex by
microvascular endothelium. J Clin Invest 92: 2224-2229.
109. Bickel U (1995) Antibody delivery through the blood-brain barrier. Adv Drug Deliv Rev 15: 53-72.
110. Bechara C, Sagan S (2013) Cell-penetrating peptides: 20 years later, where do we stand? FEBS
Lett 587: 1693-1702.
111. Jaruszewski KM, Ramakrishnan S, Poduslo JF, et al. (2012) Chitosan enhances the stability and
targeting of immuno-nanovehicles to cerebro-vascular deposits of Alzheimer's disease amyloid
protein. Nanomedicine 8: 250-260.
112. Parhamifar L, Sime W, Yudina Y, et al. (2010) Ligand-induced tyrosine phosphorylation of
cysteinyl leukotriene receptor 1 triggers internalization and signaling in intestinal epithelial cells.
PLoS One 5: e14439.
113. Malhotra M, Tomaro-Duchesneau C, Prakash S (2013) Synthesis of TAT peptide-tagged
PEGylated chitosan nanoparticles for siRNA delivery targeting neurodegenerative diseases.
Biomaterials 34: 1270-1280.
114. Sarvaiya J, Agrawal YK (2015) Chitosan as a suitable nanocarrier material for anti-Alzheimer
drug delivery. Int J Biol Macromol 72: 454-465.
115. Gao Y, Wang ZY, Zhang J, et al. (2014) RVG-peptide-linked trimethylated chitosan for delivery
of siRNA to the brain. Biomacromolecules 15: 1010-1018.
116. Sharma G, Modgil A, Layek B, et al. (2013) Cell penetrating peptide tethered bi-ligand liposomes
for delivery to brain in vivo: Biodistribution and transfection. J Control Release 167: 1-10.
© 2015 Zoya Leonenko, et al., licensee AIMS Press. This is an
open access article distributed under the terms of the Creative
Commons Attribution License
(http://creativecommons.org/licenses/by/4.0)
AIMS Molecular Science
Volume 2, Issue 3, 332-358.
|
1603.09212 | 1 | 1603 | 2016-03-30T14:28:48 | Highly Catalytic Nanodots with Renal Clearance for Radiation Protection | [
"physics.bio-ph"
] | Ionizing radiation (gamma and x-ray) is widely used in industry and medicine, but it can also pose a significant hazardous effect on health and induce cancer, physical deformity and even death, due to DNA damages and invasion of free radicals. There is therefore an urgent unmet demand in designing highly efficient radioprotectants with synergetic integration of effective renal clearance and low toxicity. In this study, we designed ultrasmall (sub-5 nm) highly catalytically active and cysteine-protected MoS2 dots as radioprotectants and investigated their application in protection against ionizing radiation. In vivo preclinical studies showed that the surviving fraction of MoS2-treated mice can appreciably increase to up to 79 % when they were exposed to high-energy ionizing radiation. Furthermore, MoS2 dots can contribute in cleaning up the accumulated free radicals within the body, repairing DNA damages and recovering all vital chemical and biochemical indicators, suggesting their unique role as free radical scavengers. MoS2 dots showed rapid and efficient urinary excretion with more than 80 % injected dose (I.D.) eliminated from the body after 24 hours due to their ultrasmall hydrodynamic size and did not cause any noticeable toxic responses up to 30 days. | physics.bio-ph | physics | Highly Catalytic Nanodots with Renal Clearance for
Radiation Protection
Xiao-Dong Zhang 1,3, *, Jinxuan Zhang 2, Junying Wang 1, Jiang Yang 4, Jie Chen 3, Xiu Shen 3,
Jiao Deng 5, Dehui Deng 5, Wei Long 3, Yuan-Ming Sun 3, Changlong Liu,1 Meixian Li 2, *
1 Department of Physics, School of Science, Tianjin University, Tianjin 300072, PR China
2 Institute of Analytical Chemistry, College of Chemistry and Molecular Engineering, Peking
University, Beijing 100871, China
3 Tianjin Key Laboratory of Molecular Nuclear Medicine, Institute of Radiation Medicine,
Chinese Academy of Medical Sciences and Peking Union Medical College, No. 238, Baidi Road,
Tianjin 300192, China
4 Environment, Energy and Natural Resources Center, Department of Environmental Science and
Engineering, Fudan University, No.220, Handan Road, 200433, China
5 State Key Laboratory of Catalysis, iChEM, Dalian Institute of Chemical Physics, Chinese
Academy of Sciences, Dalian 116023, China
Correspondence should be addressed to X.Z. ([email protected]),
M.L.([email protected])
ABSTRACT: Ionizing radiation (gamma and x-ray) is widely used in industry and medicine, but
it can also pose a significant hazardous effect on health and induce cancer, physical deformity
and even death, due to DNA damages and invasion of free radicals. There is therefore an urgent
unmet demand in designing highly efficient radioprotectants with synergetic integration of
effective renal clearance and low toxicity. In this study, we designed ultrasmall (sub-5 nm) highly
catalytically active and cysteine-protected MoS2 dots as radioprotectants and investigated their
application in protection against ionizing radiation. In vivo preclinical studies showed that the
surviving fraction of MoS2-treated mice can appreciably increase to up to 79 % when they were
exposed to high-energy ionizing radiation. Furthermore, MoS2 dots can contribute in cleaning up
the accumulated free radicals within the body, repairing DNA damages and recovering all vital
chemical and biochemical indicators, suggesting their unique role as free radical scavengers.
MoS2 dots showed rapid and efficient urinary excretion with more than 80 % injected dose (I.D.)
eliminated from the body after 24 hours due to their ultrasmall hydrodynamic size and did not
cause any noticeable toxic responses up to 30 days.
Introduction
High-energy ionizing radiations (X-rays and gamma rays) are widely used in industry and
medicine, but these radiations also cause significant health hazards such as cancer and other
related diseases.1-6 In clinical medicine, more than 50% cancer patients need to receive radiation
therapies, but high-energy radiations during the treatments not only kill cancer cells but also
cause inevitable damages to normal tissues.7-8 Besides, more and more nuclear power plants are
proposed in the world due to the rising demands in energy, imposing potential radiation risks on
public health.9-10 Under exposures to ionizing radiations, lots of free radicals including reactive
oxygen species (ROS) are formed through ionizing reactions such as the photoelectric, Compton
scattering and Auger effects.11-12 These free radicals are able to react with DNA and RNA inside
the body, cause structural and functional changes and affect biological processes.13-15 As a result,
it induces cell apoptosis and further triggers cancer or even death. The use of radioprotectants
provides a feasible solution to shield heath tissues from high-energy radiations.16 Amifostine
(Ethyol®) is an extensively used prescription radioprotectant in radiation medicine by scavenging
oxygen-derived free radicals, but its blood elimination half-life is only 1 minute, limiting its
scavenging activities against ROS.16 Other materials have shown abilities of radiation protection
to an extent, but they cannot afford effective excretions which could result in potential hepatic
and splenic toxicities.14-15 As a matter of fact, US Food and Drug Administration (FDA) have
required all injected agents for in vivo uses to be completely cleared from the body.17 Therefore,
it is highly desirable to explore an ideal radioprotectant with capabilities of highly efficient
removals of ROS, renal clearance and low toxicities, for clinical translation as an adjuvant in
radiotherapies.
To address these critical challenges, we utilized a simple approach towards the synthesis of
ultrasmall cysteine-protected MoS2 dots as radioprotectants. Surface protection with zwitterionic
cysteine offers several indispensable merits to the MoS2 dots. Firstly it avoids the significant
increase in hydrodynamic sizes as compared to other surface ligands and thus allows effective
elimination from renal clearance.18-19 Secondly, the aqueous dispersibility and stability are
significantly enhanced by the surface modification, maintaining the ultrasmall size in vivo and
refraining from aggregation. Last, non-specific adsorption of serum proteins, especially opsonin,
is largely prohibited, leading to postponed removal from the body and a relatively longer
circulation time in blood to achieve desired radiation protection. The as-prepared MoS2 dots
exhibited extraordinary electrocatalytic activities for hydrogen peroxide and oxygen reduction
reactions (ORRs), leading to lots of free electron transfers. Its endogenous catalytic properties
provide a promising and effective pathway for scavenging free radicals in vivo via rapid reactions
with oxygen radical superoxide (O2-) and non-radical oxidant hydrogen peroxide (H2O2) in blood.
Owing to their strong catalytic activities, MoS2 dots improved the surviving fraction of mice
exposed to high radiation doses of 662 keV gamma ray. Furthermore, the ultrasmall MoS2 dots
can indeed eliminate the ROS through reduction reactions in major organs. As a result,
superoxide dismutase (SOD), as an important indicator for antioxidant defense in almost all
living cells under exposure to oxygen, nearly recovered back to normal levels, suggesting the
repair of radiation-induced damages. The ultrasmall cysteine-protected MoS2 dots can be rapidly
excreted via kidney and did not cause any toxicological responses in 30 days post injection.
Results and Discussion
Ultrasmall MoS2 dots were synthesized and purified by a combinational approach of
ultrasonication and gradient centrifugation according to our published procedures.20-22 However,
it is difficult to disperse bare MoS2 dots as a stable colloid solution in water, preventing their
further applications in biological systems. A cysteine protection layer with a molecular weight of
121 Da was introduced which not only maintains the ultrasmall hydrodynamic size to meet the
cut-off size of <5.5 nm for renal clearance,17 but also endows excellent biocompatibility in
biomedical applications. Cysteine-protected MoS2 dots were obtained and stabilized through the
chemical bonding and van der Waals interactions between MoS2 and cysteine,23 and were further
purified and re-dispersed in water (Figure 1a). Obvious Tyndall effects were observed,
indicative of the existence of small particles in solution. Figure 1b exhibited a representative
image of transmission electron microscopy (TEM) which showed cysteine-protected MoS2 dots
have a core size around 2 nm by analyzing 121 dots from TEM images (Figure S1b) and a
hydrodynamic size of 3.1 nm (Figure S1c) by dynamic light scattering (DLS). X-ray
photoelectron spectroscopic (XPS) investigation of cysteine-protected MoS2 manifested Mo and
S elements from MoS2 and N element from cysteine, suggesting the formation of cysteine-MoS2
composite (Figure 1c). Additionally, the S 2p3/2 binding energy located at 164.0 eV suggested
the formation of disulfide bonds which are ascribed to the conjugation of cysteine and MoS2 dots
(Figure 1d).24 On the other hand, no signals from disulfide bonds can be found on the
corresponding counterparts, cysteine and unprotected MoS2 dots (Figure S2). Optical absorption
was investigated by UV-vis spectroscopy (Figure 1e). Cysteine alone did not show any
absorption features, while the MoS2 suspension and cysteine-protected MoS2 dots displayed
multiple absorption peaks at 393, 470, 607 and 670 nm, attributed to deep energy levels of
electronic transitions and interband excitonic transitions.25-26 Notably, the results are similar to
the absorption bands observed for MoS2 particles prepared by other methods.25-26
We evaluated the in vitro catalytic activities of cysteine-protected MoS2 dots towards
reduction of H2O2 in N2-saturated 0.01 M phosphate-buffered salin (PBS), pH = 7.4, using a
cysteine-MoS2 modified glassy carbon electrode (GCE). Figure 1f showed the cyclic
voltammetric (CV) curves of the ultrasmall cysteine-protected MoS2 dots in the presence and
absence of 5.00 mM H2O2 at the scan rate of 50 mV·s−1. Negligible reduction current was
observed in the absence of H2O2, whereas in the presence of H2O2, the reduction current
increased sharply when the scan potential shifted to more negative than -0.4 V, indicating
extraordinary catalytic activities toward reduction of H2O2 compared to voltammetric responses
of the bare GCE without modification of cysteine-protected MoS2 dots (Figure S3a). To
determine the role of cysteine-protected MoS2 dots in oxygen reduction reaction (ORR), CV
responses of cysteine-protected MoS2 dots were investigated in O2-saturated 0.01 M PBS
solution at the scan rate of 50 mV·s−1 (Figure 1g). Cysteine-protected MoS2 dots showed
electrocatalytic activity for reduction of O2 with a more positive onset potential and larger current
density compared to those of the unmodified GCE (Figure S3b). The electrochemical
measurements of cysteine-protected MoS2 dots evidenced their strong in vitro catalytic activities
in H2O2 and oxygen reduction reactions, serving as the original inspiration for us to investigate
the biological responses in radiation-injured mice.
We firstly examined the in vitro toxicities of cysteine-protected MoS2 dots in 3T3/A31 cells
using neutral red assay and MTT method which showed extremely low cytotoxicities with doses
up to 145 μg/ml (Figure S4). Viabilities of cells treated with cysteine-protected MoS2 dots in
different doses under exposures to gamma ray were significantly increased compared with those
without treatments (Figure 2a), suggesting the powerful in vitro protective behaviors of cysteine-
protected MoS2 against radiation. The DNA damages of cysteine-protected MoS2 dots with
different concentration up to 145 μg/ml also investigated by single-cell sol-gel electrophoresis,
and no significant difference on tail moments was found, suggesting negligible DNA damages
from MoS2 dots (Figure S5). DNA damages arising from radiation were further estimated
(Figure S6). No obvious cell tail moments were found in the control, but cell tails implying
DNA damages were easily identifiable with exposures to high-energy gamma ray. In contrast, the
cell tails from cells treated with cysteine-protected MoS2 dots showed distinct recoveries,
presenting effective DNA repairs. Quantitative investigations of tail moments showed that cells
treated with cysteine-protected MoS2 dots recovered to an appreciable degree from exposures of
gamma ray, as compared to cells treated only with radiation (Figure 2b).
In vivo protection from radiation using cysteine-protected MoS2 dots were investigated with
C57BL/6 mice. 200 μL cysteine-protected MoS2 dots at different concentrations were
intraperitoneally injected into mice. The mice were then exposed to high energy gamma ray
(137Cs, 3600 Ci) at the dose of 7.5 Gy. The surviving fractions of radiation-injured mice with
various injection concentrations are shown in Figure 2c. The surviving fraction of mice exposed
to high-energy radiations without injection of cysteine-protected MoS2 dots is 0 after two weeks.
This result demonstrated radiation induced acute damages and death which are consistent with
observations in previous published work.6 With injection of cysteine-protected MoS2 dots, the
surviving fractions of mice were 7.1 and 42.9 and 78.6 % at the doses of 10, 20 and 50 mg/kg.
Meanwhile, the surviving fraction of mice treated with cysteine-protected MoS2 dots (200 μL, 5
mg/ml) in the absence of radiation is 100 %, indicating low toxicities at this concentration. In
fact, cysteine protection layers were observed to show outstanding biocompatibility, in good
agreement with previous results.27-28 Besides, we also investigated the in vivo radiation protection
effects of Pt-, Co- and Ni-doped MoS2 dots as well as cysteine, but none of these metal-doped
MoS2 dots showed comparable survival rates to that of cysteine-protected MoS2 dots (Figure S7),
which may be due to the large sizes and different surface chemistries. Noticeably, a low survival
rate was observed for cysteine-treated mice, clearly providing strong evidences that the radiation
protection effect against gamma ray is from MoS2 instead of cysteine. Next, DNA damages of
irradiated mice treated with or without cysteine-protected MoS2 were assessed (Figure 2d). Total
DNA from bone marrow cells and bone marrow nucleated cells (BMNC), two dominant
indicators of ionizing radiation, were collected from control mice, irradiated mice and irradiated
mice treated with cysteine-protected MoS2. DNA from healthy mice showed an optical density
(OD) of 0.42 after 1 day, but it drastically decreased to 0.22 only after 1 day post exposure to
radiation, indicating severe DNA damages induced by radiation. In contrast, OD of total DNA for
mice treated with cysteine-protected MoS2 dots decreased to a higher value of 0.29 after 1 day.
After 7 days, OD of mice treated with cysteine-protected MoS2 dots rose to 0.38, almost
recovered to the healthy value, compared to that of 0.25 from mice with only radiation. These
results illustrate that cysteine-protected MoS2 dots can effectively decrease DNA damages from
high-energy radiations. The number of BMNC is presented as in Figure S8. Similar to the results
of DNA damages, the BMNC number decreased from 6×106 cells/mL in healthy mice to 4.3×106
cells/mL in irradiated mice after 1 day, while it was recovered to 4.6×106 cells/mL in irradiated
mice treated with cysteine-protected MoS2 dots. On the 7th day, the BMNC number in mice
treated with MoS2 dots was stably maintained at 4.6×106/mL as to a significantly decreased level
of 1.3×106/mL in mice only treated with radiation. The results clearly presented that cysteine-
protected MoS2 dots are critical in recovering the number of BMNC.
To address long-term damages caused by radiation, blood chemistry panels of irradiated
mice under 7 Gy gamma ray were studied (Figure S9). In the beginning (1 day), it is clear that
white blood cells (WBC), red blood cells (RBC) and platelets (PLT) sharply decreased after
exposure to gamma ray indicating strong radiation-induced inflammatory responses, independent
of treatments of cysteine-protected MoS2 dots. At the middle time point (7 days), WBC and PLT
were still in low levels without distinct recoveries to healthy levels, while all other indicators
started to display certain recoveries after treatments with cysteine-protected MoS2 dots. After 30
days, a large number of the indicators of mice treated with irradiation showed significant
decreases, but those receiving treatments of cysteine-protected MoS2 dots have appreciably
recovered to normal levels. Besides, radiation triggered significant changes in levels of alanine
aminotransferase (ALT) and serum creatinine (CERA) 1 and 7 days post injection (Figure S10).
On the other hand, all these indicators for MoS2-dot-treated mice were thoroughly restored to
normal 30 days post injection, while irradiated mice without MoS2 treatments presented
significant variations in the levels of ALT, albumin (ALB), blood urine nitrogen (BUN) and
globulin (GLOB). Under exposure to radiations, the acute radiations not only give rise to severe
DNA breaks and considerable decrease in the BMNC numbers, but also instantaneously elicit
inflammatory responses.29 Radiation-induced damages were the most severe after 7 days and lots
of irradiated mice even began to die. After that, the mice started to recover gradually up to the
30th day. Undoubtedly, radiation always induces certain irreparable damages of DNA and BMNC
and thus strongly affects the panels of blood chemistry and biochemistry, implying infections and
inflammations. However, the cysteine-protected MoS2 dots can free the mice from radiation-
related DNA and BMNC damages, relieve corresponding infections and thus increase the overall
surviving fractions of mice.
To further reveal the mechanism of the underlying radiation protection, superoxide
dismutase (SOD) and 3,4-Methylenedioxyamphetamine (MDA) were employed to provide some
insights. SOD are enzymes that alternately catalyze dismutation (or partitioning) of the
superoxide (O2
−) radicals into either ordinary molecular oxygen (O2) or hydrogen peroxide
(H2O2). In contrast, MDA is a harmful product generated by mice. The contents of SOD and
MDA indicate levels of damages caused by ionizing radiation. Figure 3a and b showed SOD
levels in lung and liver. When mice were exposed to gamma ray, the SOD of lung from all mice,
independent of treatments of MoS2 dots, notably decreased after 1 day. However, 7 days post
radiation, the SOD from the mice treated with cysteine-protected MoS2 dots exhibited powerful
recoveries compared to mice without treatments, clearly elucidating the increase in the ability of
ROS clearance. Figure 3c and d examined the MDA levels of lung and liver from irradiated
mice with or without treatments of cysteine-protected MoS2 dots. The MDA level in the lung of
normal mice was 8.1 nmol/mL, but it sharply increased to 22.3 nmol/mL after 1 day (Figure 3c).
In the meantime, the treatment with cysteine-protected MoS2 dots could partially shield the mice
from radiation damages with much lower MDA levels of 15.1 nmol/mL after 1 day. After 7 days,
the MDA levels decreased to 10 nmol/mL for mice with treatments of cysteine-protected MoS2
dots compared to 12.8 nmol/mL for mice only with radiation. Similarly, Figure 3d indicated the
MDA levels in the liver sharply increased from 3.6 nmol/mL to 13.2 and 12.1 nmol/mL for
treated and untreated mice, respectively. After 7 days, the mice treated with cysteine-protected
MoS2 dots recovered back to 4 nmol/mL, but untreated mice was still 5.5 nmol/mL, significant
higher than the threshold value of normal mice without irradiations. In healthy mice, SOD is
maintained at normal levels,30-31 but high-energy irradiation is well known to cause generation of
excessive amount of ROS and consumption of large quantities of SOD. Consequently, the SOD
levels from liver and lung decrease enormously. The cysteine-protected MoS2 dots can help to
get rid of the undesirable hazardous ROS in vivo and rescue the body from consuming lots of
SOD to achieve the removal of ROS and accordingly, SOD can be recovered and reserved at the
normal level. Similarly, MDA is normally present in low levels in the body. When mice suffered
from the exposure to radiation, a large quantity of MDA is produced due to the generation of
ROS in excess. The cysteine-protected MoS2 dots can decrease MDA levels of mice via
eliminating free radicals. As such, as the cysteine-protected MoS2 dots with reducing potencies
circulate into blood, they reduce the radiation-induced ROS, functioning as free radical
scavengers.
For further medical applications, pharmacokinetics and biosafety are the two most important
characteristics to be taken into consideration. We investigated the pharmacokinetics, urine
excretion and in vivo toxicities of cysteine-protected MoS2 dots at the injection dose of 50 mg/kg.
The C57/BL6 mice were injected at the concentration of 5 mg/ml with blood and urine collected
for pharmacokinetic and renal-excretion studies, as measured by inductively coupled plasma
mass spectrometry (ICP-MS). Figure 4a shows that the cysteine-protected MoS2 dots has a half
time of 2.1 hours in blood. Furthermore, nearly 80% of cysteine-protected MoS2 dots can be
rapidly excreted through the urine route after 24 hours post injection due to their ultrasmall
hydrodynamic size (Figure 4b and Figure S11). Biodistribution of cysteine-protected MoS2 dots
was analyzed after 24 hours post injection using ICP-MS. Bladder and kidney showed the highest
distribution, ascribed to renal clearance, whereas spleen, liver, lung and heart had relatively low
uptakes (Figure 4c). With increasing post injection time of up to 30 days, the uptakes of MoS2
dots from all organs sharply decreased as compared to 1 day and MoS2 dots have almost been
completely eliminated from the body. Traditional nanoparticles with larger sizes cannot be fast
excreted, inducing high uptakes in liver and spleen.32 For example, carbon nanotubes with PEG
coatings can only be excreted for a total of 60% after 90 days post injection,33-34 and lots of
nanotubes accumulated in liver and spleen will induce potential liver toxicities. Noticeably,
radiation-protective effects of CeO2 and Ag nanoparticles have been previously reported, but all
those nanomaterials are not clearable by the glomerular filtration of kidney, owing to their large
sizes as well as active surface chemistries.15, 35 The ultrasmall cysteine-protected MoS2 dots
shown here, nevertheless, provide the unique feature of highly efficient renal clearance,
resembling small molecules.
We next evaluated the in vivo toxicities of cysteine-protected MoS2 dots in terms of body
weight, immune responses, hematology and biochemistry panels at the time points of 1, 7 and 30
days. During the 30-day period, treatments with cysteine-protected MoS2 dots did not induce any
obvious adverse effects on the growth of mice and no meaningful statistical differences were
observed in the body weight and the thymus index between the cysteine-protected MoS2 dots-
treated mice and control mice. We examined standard hematological biomarkers including WBC,
RBC, hematocrit (HCT), mean corpuscular volume (MCV), hemoglobin (HGB), PLT, mean
corpuscular hemoglobin (MCH) and mean corpuscular hemoglobin concentration (MCHC).
Hematological results for the cysteine-protected MoS2 dots are presented in Figure 4d and
Figure S12a. For the mice treated with cysteine-protected MoS2 dots, two typical indicators
(WBC, RBC) did not show any meaningful differences from those of untreated mice. These
results clearly illustrate that the cysteine-protected MoS2 dots do not induce significant infections
and inflammations in mice and are relatively safe in pre-clinical settings.
In addition, we performed the standard biochemistry examination on the mice treated with
cysteine-protected MoS2 dots at different time points of 1, 7 and 30 days (Figure 4e and Figure
S12b). The biochemical parameters including ALT, AST, total protein (TP), ALB, blood urea
nitrogen, CREA, GLOB, and total bilirubin (TBIL) were investigated. We emphasize ALT, AST,
and CREA, because they are closely related to the functions of liver and kidney of mice. ALT
showed increase after 7 days, but recovered to normal after 30 days treatments. AST and CREA
did not have statistical differences in the treated mice. In fact, MoS2 nanosheets and other two-
dimensional nanomaterials have shown potential applications in imaging and photothermal
therapies, but the hydrodynamic sizes of these particles are too much larger than the 5.5 nm cut-
off value of kidney, which will induce potential liver toxicities.36-43 For example, Au
nanoparticles coated by PEG and BSA caused acute damages even after 30 days at a relatively
low injection dose of 5 mg/kg,44-45 In comparison, the sub-5 nm MoS2 dots we prepared with
efficient renal clearance will escape from the uptake of the reticuloendothelial system (ROS),
minimizing the toxic effects. The cysteine-protected MoS2 dots showed extremely low liver
toxicities even at a 10-fold high dose of 50 mg/kg. Finally, the pathological changes of organs
were demonstrated by immunohistochemistry at different time points. Liver, spleen, kidney and
other organs were collected and sliced for Haematoxylin and Eosin (H&E) staining (Figure 4f
and Figure S13). No apparent damages were observed in all organs, especially the spleen and
kidney during the entire period.
Radiation protection is utmost important in both healthcare and environment. Fukushima
nuclear radiation has caused detrimental public crisis and irreversible radiation damages in large
areas.46 Thus, the development of radioprotectants with low toxicities or even without toxicities
is promising for potential applications in this area. Cysteine-protected MoS2 dots not only endow
renal clearance, but also offer strong catalytic properties that allow to react with excessive
radiation-induced ROS in the blood of mice. However, we have conceived that it still has large
room for the development of new materials for protection from ionizing radiation. In the future,
it is valuable to explore new small-molecular organic materials with high activities in free radical
scavenging. It is necessary to develop effective radioprotectants with sub-5 nm hydrodynamic
sizes as well as low toxicities.47-54 Besides, it is of great importance to design materials with
well-defined functional surfaces for radiation protection. It is also interesting to combine features
of energy transfers and energy-dependent responses under exposure to high-energy radiation.
Detailed biological and bio-catalytic mechanisms of radiation protection are still necessary to be
further investigated.55-57
Conclusion
In summary, we developed ultrasmall cysteine-protected MoS2 dots radioprotectants, with
high efficiencies in renal clearance and radiation protection against gamma ray. This type of
novel nanomaterials behave like organic molecules with low levels of retentions in RES organs
such as liver and spleen, avoiding associated toxicities. The cysteine-protected MoS2 dots also
manifested high catalytic activities against H2O2 and O2, leading to potential removal of ROS in
vivo. Both in vitro and in vivo experiments showed that the surviving fractions of mice can be
significantly increased with exposure to high-energy gamma ray. Furthermore, treatments of
cysteine-protected MoS2 dots can considerably decrease DNA breaks and BMNC damages and
can also recover radiation-induced damages on WBC, PLT and other vital chemical and
biological indicators. The cysteine-protected MoS2 dots behave as free radical scavengers and
induce increase in SOD and decrease in MDA. Finally, we found cysteine-protected MoS2 dots
achieved around 80% urine excretion via bladder in 24 hours without any observable toxic
effects even at a high injection dose of 50 mg/kg.
Experimental Section
Preparation of Ultrasmall Cysteine-protected MoS2 Dots. Firstly MoS2 dots were
synthesized by a series of ultrasonication and centrifugation steps according to our previous
work20. MoS2 powder (99%, <2 μm in size, Aldrich) was dispersed in N,N-dimethylformamide
(DMF) (99.9%, Aldrich) at the concentration of 1 mg mL-1 and then subject to ultrasonication for
8 h using a SB-2200 sonifier (Shanghai Branson, China) to form a black suspension. The
resulting suspension was centrifuged at 6000 rpm for 30 min using a high-speed centrifuge
(Changsha Pingfan Instrument and Meter Co., Ltd., China), with the light yellow supernatant
collected. It was then centrifuged again at 12000 rpm for 30 min and the precipitated pellets were
washed with deionized water several times and redispersed in water to form an aqueous
suspension of MoS2 dots as the unprotected control. As for cysteine-protected MoS2 dots, the
precipitates were re-dispersed into DMF of the same volume as before centrifugation and L-
cysteine (99%, Beijing Chemical Reagent Co., Ltd.) was added to achieve a final concentration
of 0.1 mg mL-1. The mixture was mixed uniformly and kept still for 24 hours of aging at room
temperature. Particles in larger sizes were removed by centrifugation at 6000 rpm for 15 min and
the remaining supernatant was centrifuged again at 12000 rpm for 30 min with the precipitates
collected. The product was washed extensively with deionized water and centrifuged at 12000
rpm to remove free cysteine and residual DMF. The solution of cysteine-protected MoS2 dots was
prepared by redispersing the precipitates into certain volumes of water.
Preparation of Pt-, Co- or Ni-doped MoS2. The Pt, Co and Ni doped MoS2 (Pt-MoS2, Co-
MoS2 and Ni-MoS2) were synthesized according to our previous work.58 For Pt-MoS2, 900 mg
(NH4)6Mo7O24·4H2O and 0.442 mL 0.19 mol L-1 H2PtCl6 (aq.) were dissolved in 20 mL
deionized water to form a homogeneous solution, which were then transferred into a 40 mL
stainless steel autoclave with 10 mL CS2 under Ar and maintained at 400 oC for 4 h. The
saturated NaOH (aq.) were used to treat the obtained product at 60 oC for 3 h, followed by
washing several times with water and absolute ethanol and drying at 100 oC. The Co–MoS2 and
Ni–MoS2 were synthesized by using 900 mg (NH4)6Mo7O24·4H2O and 0.085 g Co(NO3)2·6H2O
or 0.084 g Ni(NO3)2·6H2O dissolved in 20 mL deionized water and 10mL CS2, following the
same process as for Pt–MoS2. The doping contents of Pt, Co and Ni in MoS2 were all 1.7 wt.%
measured by inductively coupled plasma optical emission spectroscopy (ICP-OES).
Modification of GCE with Ultrasmall Cysteine-protected MoS2 Dots. Cysteine-
protected MoS2 dots were redispersed in water at the concentration of 1 mg mL-1. Prior to
modification, GCE was polished with 0.3 μm alumina slurries followed by 0.05 μm alumina
slurries (Buehler). The polished GCE was sonicated three times, each for 3 min in deionized
water to remove any residual polishing reagents. The solution of MoS2 was drop casted on GCE
(diameter = 3 mm) and dried in air with a loading of 0.2 mg cm–2.
Characterization. All electrochemical measurements were carried out on a CHI 660D
electrochemical workstation (Chenhua, China) at room temperature. The cysteine- MoS2 dot-
modified GCE, a Pt electrode and a saturated calomel electrode (SCE) were used as the working,
counter and reference electrodes respectively for all electrochemical measurements. Transmission
electron microscopy (TEM) images were acquired on a JEM-2100F electron microscope (JEOL,
Japan). X-ray photoelectron spectra (XPS) were collected on an Axis Ultra spectrometer (Kratos
Analytical Ltd., Japan) and the binding energy was calibrated by the C 1s peak at 284.8 eV. UV-
vis absorption spectra were recorded with a U-4100 UV-vis-NIR spectrophotometer (Hitachi,
Japan). The hydrodynamic diameter of the cysteine-protected MoS2 nanodots was determined by
dynamic light scattering (DLS) with a NanoZS Zetasizer (Malvern). The sample solutions were
prepared by dissolving the cysteine-protected MoS2 nanodots in deionized ultrafiltered (DIUF)
water at a concentration of 0.01 mg/mL. DLS data were acquired in the phase analysis light
scattering mode at 25 °C and poly-dispersion index is always <0.7 for every measurement.
In Vitro Cytotoxicity and Radiation Protection. BALB/3T3 clone A31 mouse fibroblasts
(A31) were employed for cell viability experiments. A31 cells were dispensed in two 96-well
plates with 7x103 cells per well and incubated under 37℃, 5% CO2 and humidified atmosphere.
Different concentrations of MoS2 from 0.5-135 μg/ml were introduced into the DMEM culture
media. After incubation for 24 and 48 hours, cytotoxicities were analyzed with Neutral Red (NR)
staining. Meanwhile, to further confirm the accuracy of the cell survival rates, cysteine-protected
MoS2 nanodots were also assessed using MTT Cell Proliferation and Cytotoxicity Assay Kit.
Cysteine-protected MoS2 dots (final concentrations of 0.5-135 μg/ml) were introduced into the
DMEM culture media. After 24 or 48 h of treatments, 10 μL of MTT reagent was added to each
well and incubated for 4 h, and then the media were replaced with 150 μL DMSO to dissolve
formazan crystals. The optical absorption in 490 nm was measured with a single tube
luminometer (TD 20/20, Turner Biosystems Inc., Sunnyvale, CA, USA) with the intrinsic optical
absorption of cysteine-protected MoS2 nanodots subtracted. The cell survival rates were
calculated based on the recorded optical absorption.
4x103 A31 cells were seeded into five 96-well plates, supplemented with high-glucose
DMEM containing 10% FBS. After 10 hours of incubation, the cells were treated with 0.2 μg and
2 μg MoS2 dots for 1 hour which were dissolved in 100 ul complete culture media. The five
plates were then exposed to 0, 2, 4, 8 and 10 Gy radiations respectively. After 48 hours, NR was
used to evaluate cell viabilities. When cells had sufficiently taken up NR in about~3 hours, the
media containing NR was discarded and each well was gently washed with 1X PBS, followed by
addition of 150 μl desorb solution to extract NR from living cells. All the plates were
homogenized on a microtiter plate shaker for 5 min and the optical absorbance was recorded at
540 nm by Infinite F200 multimode plate reader.
In Vitro DNA Damages. The in vitro protocols were approved by the Institutional Review
Board of Institute of Radiation Medicine at the Chinese Academy of Medical Sciences (CAMS).
A modified protocol of alkaline COMET assay was used to evaluate DNA damages. Agarose
(0.8%, 500 μl) was paved homogeneously on glass microscopic slides. After solidification, 4.8
x104 cells in 30 μL 1X PBS incubated with MoS2 dots (0.5-135 μg/ml) were mixed with 70 μL
low-melting-point agarose (0.6%) and 20 μl of this mixture was spread over the slide completely.
The solidified slides were placed into freshly-prepared ice-cold lysis buffer (2.5 mol/L NaCl, 100
mmol/L Na2EDTA, 10 mmol/L Tris-HCl, 10% DMSO, 1% Triton X-100) for 2 hours, followed
by immersing them in a horizontal gel electrophoresis unit which was filled with chilled
electrophoresis buffer (1 mmol/L Na2EDTA, 300 mmol/L NaOH, pH 7.4 for 30 min.
Electrophoresis was conducted at 30 V for 20 min. The slides were neutralized with ethanol after
electrophoresis and stained with ethidium bromide (2 μg/mL). DNA damages were analyzed by
Comet Assay Software Project (CASP) for tail moments. Typically, 100 cells were analyzed for
DNA tail moment.
In Vivo Radiation Protection. All animal-related protocols were reviewed and approved by
the Institutional Animal Care and Use Committee (IACUC). All animals were purchased,
maintained and handled under protocols approved by the Institute of Radiation Medicine at the
Chinese Academy of Medical Sciences (CAMS). 200 μL saline was intraperitoneally injected
into mice as the control group. As for the treatment group, 1, 2 or 5 mg/mL cysteine-protected
MoS2 dots were used for animal experiments by intraperitoneal injection with doses of 10, 20
and 50 mg/kg in mice. Subsequently, the mice were radiated under 7.5 Gy gamma rays (~8 min)
and 137Cs with an activity of 3600 Ci and a photon energy of 662 KeV were used. 70 mice in total
were assigned into the following 5 groups (14 mice in each group): control, only treated with
radiation, treated with both radiation and cysteine-protected MoS2 dots at the doses of 10, 20 and
50 mg/kg. We then monitored all the mice for up to 30 days and the surviving fractions of mice
were obtained for each group.
Analysis of Total DNA and Bone Marrow Nucleated Cells: 32 male C57BL/6 mice were
assigned into the following 4 groups (8 mice each group): control of 1 day, treatment group of 1
day, control of 8 days and treatment group of 8 days. For the control groups, mice were
administered with 200 μL distilled water and the treatment groups were intraperitoneally injected
with 200 μl solution of cysteine-protected MoS2 dots at the dose of 50 mg/kg. The entire body of
each mouse was radiated under gamma ray at the dose of 6.5 Gy. After 1 and 8 days of treatments,
mice were sacrificed. Bilateral femurs of each mouse were excised from the body and connective
tissues were removed completely. To estimate the total DNA levels in bone marrows, bone
marrow cells were flushed from the femurs into calcium chloride solution (10ml, 5 mM) with a
24-gauge needle and as such, single-cell suspensions were made. The suspensions were placed at
4℃ for 30 minutes and centrifuged at 2500 rpm/min for 15 minutes with supernatants discarded.
The pellets were mixed with perchlorate (5 mL 0.2 M) and stored in water bath being heated for
15 min at 90℃. After being cooled down to room temperature, the mixtures were purified
through filter paper (pore size =0.2 μm) and the filtrates were measured for UV-vis absorption at
268 nm (Shimadzu, UV-1750). Similarly, bone marrow nucleated cells were flushed into 1 mL
1X PBS, filtered through nylon meshes to remove fragments of bones and tissues for cytometric
analysis (Mindray BC-2800 Vet).
SOD and MDA Analysis: All the organs for SOD and MDA analysis were from identically
treated mice as for total DNA and BMNC analysis. After 1 and 8 days of treatments, mice were
sacrificed with livers and lungs collected for analysis of SOD and MDA levels using Total
Superoxide Dismutase assay kits and Malondialdehyde assay kits (Nanjing Jiancheng
Bioengineering Institute). All the organs were immersed in saline solution and homogenized with
a tissue homogenate machine (IKA, T18 basic) to make 10% tissue homogenate. The organ
homogenates were left on ice for 1 hour and then centrifuged at 3500 r/min for 10 minutes. The
resulting solutions were further diluted into 1 % and 0.25 % homogenate respectively. For total
SOD level analysis: phosphate buffer (1 ml, 7.5 mM) was added to 5 ml EP tubes and then mixed
with liver and lung homogenate (50 μl, 0.25 %) or 50 μl distilled water as control.
Hydroxylamine hydrochloride (0.1 M, 100 μl), xanthine solution (75 mM, 100 μl) and
xanthineoxidase solution (0.03 U/L, 100 μl) were added into the above EP tubes in order, fully
mixed and incubated at 37 ℃ for 40 minutes. After that, 2ml nitrite developer (C10H9N:
C6H7NO3S: CH3COOH=3:3:2) was added
into
the system and analyzed by UV-vis
spectrophotometer (Shimadzu, UV-1750) at 550 nm. For MDA analysis, 10% tissue
homogenates were employed. Equal volumes (100 μl) of organ sample, ethanol and
tetrathoxypropane (10 nmol/ml) were added into individual 5 ml EP tubes, among which ethanol
and tetrathoxypropane were negative and positive controls respectively. 100 ul tissue lysis
solution was added into each 5ml EP tube and mixed completely. Trichloroacetic acid (10 %, 3
ml) and thiobarbituric acid (0.6 %, 1 ml) were added into the system orderly. Then the tubes were
heated in water at 95℃ for 40 minutes. After being cooled down to room temperature, all the
samples were centrifuged at 3500~4000 r/min for 10 minutes with the supernatants analyzed by
UV-vis spectrophotometer at 532 nm (Shimadzu, UV-1750).
In Vivo Toxicity. Animals were purchased, maintained, and handled with protocols
approved by the Institute of Radiation Medicine, Chinese Academy of Medical Sciences (IRM,
CAMS). 48 male C57BL/6 mice at the age of 11 weeks were obtained from IRM laboratories,
housed by 2 mice per cage with a 12 hours/12 hours light/dark cycle and were provided with food
and water. Mice were randomly divided into 6 groups (8 mice in each group): control of 1 day,
treatment group of 1 day, control of 7 days, treatment group of 7 days, control of 30 days, and
treatment group of 30 days respectively. The solution of cysteine-protected MoS2 dots (200 μL, 5
mg/mL) was used for this animal experiment through intraperitoneal injection. The assigned dose
was 50 mg/kg in each mouse. Mice were weighed and assessed for behavioral changes every day
post injection. After treatments with cysteine-protected MoS2 dots for 1, 7 and 30 days, mice
were sacrificed accordingly at each time point using isoflurane anesthetic and angiocatheter
exsanguination with 1X PBS. Blood and organs were collected for biochemistry and pathological
analyses. One mouse from each group was fixed with 10% formalin in PBS following
exsanguination with PBS. During necropsy, liver, kidneys, spleen, heart, lung, testis, brain,
bladder and thymus were collected and weighed.
Hematology, Biochemistry and Pathology: Using a standard saphenous technique, vein
blood was collected in potassium EDTA collection tubes for hematological analysis. Standard
hematological and biochemical examinations were performed. 1 mL of blood was collected from
mice and separated into cellular and plasma fractions by centrifugation. Mice were sacrificed by
isoflurane anesthetic and angio catheter exsanguinations. Major organs were harvested, fixed in
10% neutral buffered formalin, processed routinely into paraffin and stained by H&E, with
pathology examined using a digital microscope.
Statistical Analyses: Paired Student's t-test was used for statistical analysis between the
control group and the treatment group (MoS2 alone or radiation+MoS2). The Analysis of
Variance (ANOVA) was used for statistical analysis among the control, radiation, and
radiation+MoS2 groups. Differences were considered statistically significant at a P value < 0.05.
ACKNOWLEDGMENT
This work was supported by the National Natural Science Foundation of China (Grant
No.81471786 and 21475003 and 21275010), Natural Science Foundation of Tianjin (Grant No.
13JCQNJC13500).
Author contributions.
X.Z. and M.L, conceived and designed the experiments. J.W., J.Z., W.L., and J.C., X.S., J.D.,
performed the all experiments. J.W., X.S., and W.L., contributed to radiation protection. J.Z.,
and .J.D, contributed to synthesis. X.Z., M.L, J.Y, J.W., J.Z., W.L., J.C., X.S., J.D., D. D., and
Y.S. analyzed the data and wrote the manuscript. All authors discussed the results and
commented on the manuscript.
REFERENCES
Brenner, D. J.; Elliston, C. D.; Hall, E. J.; Berdon, W. E., Estimated Risks of Radiation-
1.
Induced Fatal Cancer from Pediatric CT. Am. J. Roentgenol. 2001, 176 (2), 289-296.
2.
Brenner, D. J.; Doll, R.; Goodhead, D. T.; Hall, E. J.; Land, C. E.; Little, J. B.; Lubin, J.
H.; Preston, D. L.; Preston, R. J.; Puskin, J. S., Cancer Risks Attributable to Low Doses of
Rossi, H. H.; Kellerer, A. M., Radiation Carcinogenesis at Low Doses. Science 1972,
Newhauser, W. D.; Durante, M., Assessing the Risk of Second Malignancies after
Shirazi, A.; Ghobadi, G.; Ghazi-Khansari, M., A Radiobiological Review on Melatonin: a
Peters, E.; Slovic, P., The Role of Affect and Worldviews as Orienting Dispositions in the
Rhodes, R.; Beller, D., The Need for Nuclear Power. Foreign Affairs. 2000, 79 (1), 30-
Hainfeld, J. F.; Dilmanian, F. A.; Slatkin, D. N.; Smilowitz, H. M., Radiotherapy
Seymour, C. B.; Mothersill, C., Radiation-Induced Bystander Effects-Implications for
Ionizing Radiation: Assessing What We Really Know. P. Nati. Acad. Sci. USA. 2003, 100 (24),
13761-13766.
3.
Chen, H.; Wang, G. D.; Chuang, Y.-J.; Zhen, Z.; Chen, X.; Biddinger, P.; Hao, Z.; Liu,
F.; Shen, B.; Pan, Z., Nanoscintillator-Mediated X-ray Inducible Photodynamic Therapy for in
Vivo Cancer Treatment. Nano Lett. 2015, 15 (4), 2249-2256.
4.
Cancer. Nat. Rev. Cancer 2004, 4 (2), 158-164.
5.
(4018), 200-202.
6.
Fan, S.; Meng, Q.; Xu, J.; Jiao, Y.; Zhao, L.; Zhang, X.; Sarkar, F. H.; Brown, M. L.;
Dritschilo, A.; Rosen, E. M., DIM (3, 3′-diindolylmethane) Confers Protection against Ionizing
Radiation by a Unique Mechanism. P. Nati. Acad. Sci. USA. 2013, 110 (46), 18650-18655.
7.
Modern Radiotherapy. Nat. Rev. Cancer 2011, 11 (6), 438-448.
8.
Novel Radioprotector. J. Radiat. Res. (Tokyo) 2007, 48 (4), 263-272.
9.
Perception and Acceptance of Nuclear Power. J. Appl. Soc. Psychol. 1996, 26 (16), 1427-1453.
10.
44.
11.
Enhancement with Gold Nanoparticles. J. Pharm. Pharmacol. 2008, 60 (8), 977-986.
12.
Devasagayam, T.; Tilak, J.; Boloor, K.; Sane, K. S.; Ghaskadbi, S. S.; Lele, R., Free
Radicals and Antioxidants in Human Health: Current Status and Future Prospects. Japi 2004, 52
(794804), 4.
Colon, J.; Hsieh, N.; Ferguson, A.; Kupelian, P.; Seal, S.; Jenkins, D. W.; Baker, C. H.,
13.
Cerium Oxide Nanoparticles Protect Gastrointestinal Epithelium from Radiation-Induced
Damage by Reduction of Reactive Oxygen Species and Upregulation of Superoxide Dismutase 2.
Nanomedicine -Nanotechnol. 2010, 6 (5), 698-705.
14.
for Protection from Radiation-Induced Cellular Damage. Nano Lett. 2005, 5 (12), 2573-2577.
15.
Colon, J.; Herrera, L.; Smith, J.; Patil, S.; Komanski, C.; Kupelian, P.; Seal, S.; Jenkins,
D. W.; Baker, C. H., Protection from Radiation-Induced Pneumonitis using Cerium Oxide
Nanoparticles. Nanomedicine -Nanotechnol. 2009, 5 (2), 225-231.
16.
Directions. Oncology 2002, 63 (2), 2-10.
17.
Frangioni, J. V., Renal Clearance of Quantum Dots. Nat. Biotechnol. 2007, 25 (10), 1165-1170.
18.
Efficient Renal Clearance. Angew. Chem. Int. Ed. 2011, 123 (14), 3226-3230.
19.
Chou, L. Y.; Zagorovsky, K.; Chan, W. C., DNA Assembly of Nanoparticle
Superstructures for Controlled Biological Delivery and Elimination. Nat. Nanotechnol. 2014, 9
(2), 148-155.
20. Wang, T. Y.; Liu, L.; Zhu, Z. W.; Papakonstantinou, P.; Hu, J. B.; Liu, H. Y.; Li, M. X.,
Enhanced Electrocatalytic Activity for Hydrogen Evolution Reaction from Self-Assembled
Zhou, C.; Long, M.; Qin, Y.; Sun, X.; Zheng, J., Luminescent Gold Nanoparticles with
Grdina, D. J.; Murley, J. S.; Kataoka, Y., Radioprotectants: Current Status and New
Choi, H. S.; Liu, W.; Misra, P.; Tanaka, E.; Zimmer, J. P.; Ipe, B. I.; Bawendi, M. G.;
Tarnuzzer, R. W.; Colon, J.; Patil, S.; Seal, S., Vacancy Engineered Ceria Nanostructures
Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.-Y.; Galli, G.; Wang, F.,
Monodispersed Molybdenum Sulfide Nanoparticles on an Au Electrode. Energ. Environ. Sci.
2013, 6 (2), 625-633.
21. Wang, T.; Zhu, H.; Zhuo, J.; Zhu, Z.; Papakonstantinou, P.; Lubarsky, G.; Lin, J.; Li, M.,
Biosensor Based on Ultrasmall MoS2 Nanoparticles for Electrochemical Detection of H2O2
Released by Cells at the Nanomolar Level. Anal. Chem. 2013, 85 (21), 10289-10295.
22. Wang, T. Y.; Gao, D. L.; Zhuo, J. Q.; Zhu, Z. W.; Papakonstantinou, P.; Li, Y.; Li, M. X.,
Size-Dependent Enhancement of Electrocatalytic Oxygen-Reduction and Hydrogen-Evolution
Performance of MoS2 Particles. Chem. Eur. J. 2013, 19 (36), 11939-11948.
23. Moehl, T.; Abd El Halim, M.; Tributsch, H., Photoelectrochemical studies on the n-
MoS2-Cysteine interaction. J. Appl. Electrochem. 2006, 36 (12), 1341-1346.
24.
Cavalleri, O.; Gonella, G.; Terreni, S.; Vignolo, M.; Floreano, L.; Morgante, A.; Canepa,
M.; Rolandi, R., High resolution X-ray photoelectron spectroscopy of L-cysteine self-assembled
films. Phys. Chem. Chem. Phys. 2004, 6 (15), 4042-4046.
25.
Emerging photoluminescence in monolayer MoS2. Nano Lett. 2010, 10 (4), 1271-1275.
26. Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F., Atomically thin MoS2: a New
Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, 105 (13), 136805.
27. Wei, Z.; Sun, L.; Liu, J.; Zhang, J. Z.; Yang, H.; Yang, Y.; Shi, L., Cysteine Modified
Rare-Earth Up-converting Nanoparticles for in Vitro and in Vivo Bioimaging. Biomaterials
2014, 35 (1), 387-392.
28. Wang, F.; Liu, X.; Lu, C.-H.; Willner, I., Cysteine-Mediated Aggregation of Au
Nanoparticles: the Development of a H2O2 Sensor and Oxidase-Based Biosensors. ACS Nano
2013, 7 (8), 7278-7286.
29.
Roots, R.; Okada, S., Protection of DNA Molecules of Cultured Mammalian Cells from
Radiation-Induced Single-Strand Scissions by Various Alcohols and SH Compounds. Int. J.
Radiat. Biol. Relat. Stud. Phys. Chem. Med. 1972, 21 (4), 329-342.
30.
Lee, J. H.; Choi, I. Y.; Kil, I. S.; Kim, S. Y.; Yang, E. S.; Park, J.-W., Protective Role of
Superoxide Sismutases against Ionizing Radiation in Yeast. BBA-Gen. Subjects 2001, 1526 (2),
191-198.
31.
Cellular Damage Caused by Ionizing Radiation. Proc. Soc. Exp. Biol. Med. 2000, 225 (1), 9-22.
32.
Liu, Y.; Chen, S.; Zhong, L.; Wu, G., Preparation of High-Stable Silver Nanoparticle
Dispersion by Using Sodium Alginate as a Stabilizer under Gamma Radiation. Radiat. Phys.
Chem. 2009, 78 (4), 251-255.
33.
Liu, Z.; Davis, C.; Cai, W.; He, L.; Chen, X.; Dai, H., Circulation and Long-term Fate of
Functionalized, Biocompatible Single-walled Carbon Nanotubes in Mice Probed by Raman
Spectroscopy. P. Nati. Acad. Sci. USA 2008, 105 (5), 1410-1415.
34.
Schipper, M. L.; Nakayama-Ratchford, N.; Davis, C. R.; Kam, N. W. S.; Chu, P.; Liu, Z.;
Sun, X.; Dai, H.; Gambhir, S. S., A Pilot Toxicology Study of Single-walled Carbon Nanotubes
in a Small Sample of Mice. Nat. Nanotechnol. 2008, 3 (4), 216-221.
Liu, Y.; Chen, S.; Zhong, L.; Wu, G., Preparation of High-stable Silver Nanoparticle
35.
Dispersion by Using Sodium Alginate as a Stabilizer Under Gamma Radiation. Radiat. Phys.
Chem. 2009, 78 (4), 251-255.
36.
Liu, T.; Wang, C.; Gu, X.; Gong, H.; Cheng, L.; Shi, X.; Feng, L.; Sun, B.; Liu, Z., Drug
Delivery with PEGylated MoS2 Nanosheets for Combined Photothermal and Chemotherapy of
Cancer. Adv. Mater. 2014, 26 (21), 3433-3440.
Karbownik, M.; Reiter, R. J., Antioxidative Effects of Melatonin in Protection against
Fan, W.; Bu, W.; Zhang, Z.; Shen, B.; Zhang, H.; He, Q.; Ni, D.; Cui, Z.; Zhao, K.; Bu, J.,
for On-Demand Depth-Independent Hypoxic
Yin, W.; Yan, L.; Yu, J.; Tian, G.; Zhou, L.; Zheng, X.; Zhang, X.; Yong, Y.; Li, J.; Gu,
37.
Z., High-Throughput Synthesis of Single-Layer MoS2 Nanosheets as a Near-Infrared
Photothermal-Triggered Drug Delivery for Effective Cancer Therapy. ACS Nano 2014, 8 (7),
6922-6933.
38.
Liu, J.; Zheng, X.; Yan, L.; Zhou, L.; Tian, G.; Yin, W.; Wang, L.; Liu, Y.; Hu, Z.; Gu, Z.,
Bismuth Sulfide Nanorods as a Precision Nanomedicine for in Vivo Multimodal Imaging-Guided
Photothermal Therapy of Tumor. ACS Nano 2015, 9 (1), 696-707.
39. Zhang, X. D.; Chen, J.; Min, Y.; Park, G. B.; Shen, X.; Song, S. S.; Sun, Y. M.; Wang, H.;
Long, W.; Xie, J., Metabolizable Bi2Se3 Nanoplates: Biodistribution, Toxicity, and Uses for
Cancer Radiation Therapy and Imaging. Adv. Funct. Mater. 2014, 24 (12), 1718-1729.
40.
X-ray Radiation-Controlled NO-Release
Radiosensitization. Angew. Chem. Int. Ed. 2015, 127 (47), 14232-14236.
41. Ma, M.; Huang, Y.; Chen, H.; Jia, X.; Wang, S.; Wang, Z.; Shi, J., Bi2S3-Embedded
Mesoporous Silica Nanoparticles for Efficient Drug Delivery and Interstitial Radiotherapy
Sensitization. Biomaterials 2015, 37, 447-455.
42. Wang, S.; Li, X.; Chen, Y.; Cai, X.; Yao, H.; Gao, W.; Zheng, Y.; An, X.; Shi, J.; Chen,
H., A Facile One-Pot Synthesis of a Two-Dimensional MoS2/Bi2S3 Composite Theranostic
Nanosystem for Multi-Modality Tumor Imaging and Therapy. Adv. Mater. 2015, 27 (17), 2775-
2782.
43. Wang, S.; Chen, Y.; Li, X.; Gao, W.; Zhang, L.; Liu, J.; Zheng, Y.; Chen, H.; Shi, J.,
Injectable 2D MoS2-Integrated Drug Delivering Implant for Highly Efficient NIR-Triggered
Synergistic Tumor Hyperthermia. Adv. Mater. 2015, 27 (44), 7117-7122.
44.
Zhang, X. D.; Wu, D.; Shen, X.; Liu, P. X.; Fan, F. Y.; Fan, S. J., In Vivo Renal
Clearance, Biodistribution, Toxicity of Gold Nanoclusters. Biomaterials 2012, 33 (18), 4628-
4638.
45.
Zhang, X. D.; Wu, D.; Shen, X.; Chen, J.; Sun, Y. M.; Liu, P. X.; Liang, X. J., Size-Dependent
Radiosensitization of PEG-Coated Gold Nanoparticles for Cancer Radiation Therapy.
Biomaterials 2012, 33 (27), 6408-6419.
46.
Plants on Marine Radioactivity. Environ. Sci. Technol. 2011, 45 (23), 9931-9935.
47.
Liu, F.; He, X.; Chen, H.; Zhang, J.; Zhang, H.; Wang, Z., Gram-Scale Synthesis of
Coordination Polymer Nanodots with Renal Clearance Properties for Cancer Theranostic
applications. Nat. Commun. 2015, 6, 8003.
48.
Zhou, M.; Li, J.; Liang, S.; Sood, A. K.; Liang, D.; Li, C., CuS Nanodots with Ultrahigh
Efficient Renal Clearance for Positron Emission Tomography Imaging and Image-Guided
Photothermal Therapy. ACS Nano 2015, 9 (7), 7085-7096.
49. Chen, H.; Wang, G. D.; Tang, W.; Todd, T.; Zhen, Z.; Tsang, C.; Hekmatyar, K.; Cowger,
T.; Hubbard, R. B.; Zhang, W., Gd-Encapsulated Carbonaceous Dots with Efficient Renal
Clearance for Magnetic Resonance Imaging. Adv. Mater. 2014, 26 (39), 6761-6766.
50.
Efficient Near-Infrared Photothermal Cancer Therapy. Small 2014, 10 (15), 3139-3144.
51.
Zhang, X. D.; Luo, Z.; Chen, J.; Shen, X.; Song, S.; Sun, Y.; Fan, S.; Fan, F.; Leong, D.
T.; Xie, J., Ultrasmall Au10−12(SG)10−12 Nanomolecules for High Tumor Specificity and Cancer
Radiotherapy. Adv. Mater. 2014, 26 (26), 4565-4568.
Tang, S.; Chen, M.; Zheng, N., Sub-10-nm Pd Nanosheets with Renal Clearance for
Buesseler, K.; Aoyama, M.; Fukasawa, M., Impacts of the Fukushima Nuclear Power
Yu, M.; Liu, J.; Ning, X.; Zheng, J., High-contrast Noninvasive Imaging of Kidney
52.
Clearance Kinetics Enabled by Renal Clearable Nanofluorophores. Angew. Chem. Int. Ed. 2015,
54 (51), 15434-15438.
53.
Zhang, X. D.; Luo, Z.; Chen, J.; Wang, H.; Song, S. S.; Shen, X.; Long, W.; Sun, Y. M.;
Fan, S.; Zheng, K., Storage of Gold Nanoclusters in Muscle Leads to their Biphasic in Vivo
Clearance. Small 2015, 11 (14), 1683-1690.
54.
Liu, Y.; Chen, C.; Qian, P.; Lu, X.; Sun, B.; Zhang, X.; Wang, L.; Gao, X.; Li, H.; Chen,
Z., Gd-Metallofullerenol Nanomaterial as Non-Toxic Breast Cancer Stem Cell-Specific Inhibitor.
Nat. Commun. 2015, 6.
55. Wang, L.; Sun, Q.; Wang, X.; Wen, T.; Yin, J. J.; Wang, P.; Bai, R.; Zhang, X. Q.; Zhang,
L. H.; Lu, A. H., Using Hollow Carbon Nanospheres as a Light-Induced Free Radical Generator
to Overcome Chemotherapy Resistance. J. Am. Chem. Soc. 2015, 137 (5), 1947-1955.
56.
Schug, H.; Isaacson, C. W.; Sigg, L.; Ammann, A. A.; Schirmer, K., Effect of TiO2
Nanoparticles and UV Radiation on Extracellular Enzyme Activity of Intact Heterotrophic
Biofilms. Environ. Sci. Technol. 2014, 48 (19), 11620-11628.
57.
Li, H.; Yang, Z. Y.; Liu, C.; Zeng, Y. P.; Hao, Y. H.; Gu, Y.; Wang, W. D.; Li, R.,
PEGylated Ceria Nanoparticles used for Radioprotection on Human Liver Cells under γ-ray
Irradiation. Free Radic. Biol. Med. 2015, 87, 26-35.
Deng, J.; Li, H.; Xiao, J.; Tu, Y.; Deng, D.; Yang, H.; Tian, H.; Li, J.; Ren, P.; Bao, X.,
58.
Triggering the Electrocatalytic Hydrogen Evolution Activity of the Inert Two-dimensional MoS2
Surface via Single-atom Metal Doping. Energy Environ. Sci. 2015, 8 (5), 1594-1601.
Figures and Figure Captions
Schematic illustration of radiation protection with cysteine-protected MoS2 dots
Figure 1 Characterization and electrocatalytic properties of cysteine-protected MoS2 dots.
(a) Schematic preparation of cysteine-protected MoS2 dots. (b) TEM image of a population of
cysteine-protected MoS2 dots with a homogeneous distribution of around 2 nm. Scale bar, 20 nm.
(c) Wide survey X-ray photoelectron spectrum and (d) S 2p spectrum of cysteine-protected MoS2
dots. The presence of disulfide bonds indicates the bonding between cysteine and MoS2. Arrows
show corresponding electronic states in MoS2 dots. (e) UV-vis spectra of aqueous solutions of
cysteine, unprotected and cysteine-protected MoS2 dots. (f) CVs of GCE modified with cysteine-
protected MoS2 dots in the presence (dotted) and absence (solid) of 5.00 mM H2O2 in N2-
saturated 0.01 M pH 7.4 PBS. Scan rates: 50 mV s–1. (g) CVs of GCE modified with cysteine-
protected MoS2 dots in N2- (solid) and O2-saturated (dotted) 0.01 M pH 7.4 PBS. Scan rate: 50
mV s−1.
Figure 2 Radiation protection in vitro and in vivo. (a) Radiation dose-dependent protection in
vitro with different injected doses (50 and 100 μg/mL) or without treatments of cysteine-
protected MoS2 dots. The survival rates of A31 cells with treatments of cysteine-protected MoS2
dots increased with increasing radiation doses, implying a significant function of protection
against ionizing radiation in vitro. (b) Cell tail moment of mice with or without treatments of
cysteine-protected MoS2 dots, suggesting of its role in DNA repairs. Tail moment of DNA
fragments from single cells receiving 4 Gy radiation was quantitatively calculated by counting
100 individual healthy and irradiated cells respectively and significant DNA recovery could be
observed after being treated with cysteine-protected MoS2 dots. (c) Survival curves of mice with
different doses (10, 20 and 50 mg/kg) or without treatments of cysteine-protected MoS2 dots
(n=14 mice/group) showing an overall 79% survival proportion after being treated with a dose of
5 mg/ml. (d) DNA damages of mice 1 and 7 days after treatments of cysteine-protected MoS2
dots (n=14 mice/group), as measured by UV-vis absorption at 268 nm. *P<0.05 as compared
with the control group (paired Student's t-test).
Figure 3 Mechanisms of protection against radiation. Superoxide dismutase (SOD) levels in
(a) lung and (b) liver 1 and 7 days after gamma radiation with or without treatments of cysteine-
protected MoS2 dots (n=10 mice/group). 3,4-Methylenedioxyamphetamine (MDA) levels in (c)
lung and (d) liver 1 and 7 days after gamma radiation with or without treatments of cysteine-
protected MoS2 dots (n=8 mice/group). Both SOD and MDA of the lung and liver recovered to
healthy levels 7 days after treatments. *P<0.05 as compared with the control group (Student's t-
test).
Figure 4 In vivo pharmacokinetics, renal clearance, biodistribution and toxicities. (a) Time-
dependent concentrations in blood after intraperitoneal injection (b) Cumulative urine excretion
at different time points. A large proportion of cysteine-protected MoS2 dots in blood were cleared
through the renal route. (c) Biodistribution of cysteine-protected MoS2 dots at 24 hours and 30
days post injection. %ID/g=percentage of the injected dose per gram of weight. (d)
Hematological data of RBC and WBC in mice treated with cysteine-protected MoS2 dots. Data
were collected at different time points of 1, 7 and 30 days after intraperitoneal injection (50
mg/kg). (e) Blood biochemistry analysis of mice treated with the cysteine-protected MoS2 dots at
different time points of 1, 7 and 30 days after treatments. *P<0.05 as compared with the control
group (Student's t-test). (f) Pathological evaluation of liver, spleen and kidney of mice treated
with cysteine-protected MoS2 dots. (n=8 mice/group) collect at different time points of 1, 7 and
30 days post injection. Scale bars, 50 μm.
Supporting Information
Highly Catalytic Nanodots with Renal Clearance for
Radiation Protection
Xiao-Dong Zhang 1,3, *, Jinxuan Zhang 2, Junying Wang 1, Jiang Yang 4, Jie Chen 3, Xiu Shen 3,
Jiao Deng 5, Dehui Deng 5, Wei Long 3, Yuan-Ming Sun 3, Changlong Liu,1 Meixian Li 2, *
1 Department of Physics, School of Science, Tianjin University, Tianjin 300072, PR China
2 Institute of Analytical Chemistry, College of Chemistry and Molecular Engineering, Peking
University, Beijing 100871, China
3 Tianjin Key Laboratory of Molecular Nuclear Medicine, Institute of Radiation Medicine,
Chinese Academy of Medical Sciences and Peking Union Medical College, No. 238, Baidi Road,
Tianjin 300192, China
4 Environment, Energy and Natural Resources Center, Department of Environmental Science and
Engineering, Fudan University, No.220, Handan Road, 200433, China
5 State Key Laboratory of Catalysis, iChEM, Dalian Institute of Chemical Physics, Chinese
Academy of Sciences, Dalian 116023, China
Correspondence should be addressed to X.Z. ([email protected]),
M.L.([email protected])
Figure S1 (a) Schematic structure (not to scale) and (b) core size distribution of cysteine-
protected MoS2 dots measured by analyzing 121 nanodots from TEM images. (c) Hydrodynamic
size of cysteine-protected MoS2 dots measured by DLS.
Figure S2 XPS spectra of S 2p region for (a) unprotected MoS2 dots and (b) cysteine. No
disulfide states could be observed in both cases.
.
Figure S3 Electrocatalytic activities of unmodified GCE. (a) CVs of unmodified GCE with
(dotted) or without (solid) addition of 5.00 mM H2O2 in N2-saturated 0.01 M pH 7.4 PBS, (b)
CVs of unmodified GCE in N2- (solid) and O2-saturated (dotted) 0.01 M pH 7.4 PBS. Scan rate:
50 mV s−1. The electrocatalytic measurements indicate low reduction activities of unmodified
GCE against H2O2 and O2 without modification of cysteine protected-MoS2 dots. Note that the
as-shown current densities are in a much smaller scale than in Figure. 1.
Figure S4 Evaluation of cytotoxicities on cysteine-protected MoS2 dots. In vitro cytotoxicities
of cysteine-protected MoS2 dots on A31 mouse fibroblasts estimated by (a) Neutral Red assays
and (b) MTT assays. The results were consistent and demonstrated low cytotoxicities of cysteine-
protected MoS2 dots.
Figure S5 Evaluation of DNA damages on cysteine-protected MoS2 dots. In vitro images of
comet assays on A31 cells treated with different concentrations of MoS2 dots (0.5-135 μg/ml)
and (h) corresponding tail moment analysis.
Figure S6 (a-c) Images of untreated control A31 cells, cells treated with only radiation and
cysteine-protected MoS2 dots plus radiation. The scale bar represent 100 μm.
Figure S7 (a) Schematic diagram showing Pt-, Co- or Ni-doped MoS2 dots. (b) Hydrodynamic
sizes of different metal-doped MoS2 dots. (c) Body weight and (d) survival curves of mice
treated with Pt-, Co-, Ni-doped MoS2 dots and cysteine at a dose of 50 mg/kg under exposure to
gamma ray. Compared with cysteine-protected MoS2 dots, neither metal doping in MoS2 dots nor
pure cysteine showed increased survival proportions or reservation in body weight (n=10
mice/group).
Figure S8 Effects on BMNC. Counts of bone marrow nucleated cells in control mice, mice only
treated with radiation and mice treated with radiation and cysteine protected-MoS2 dots at the
dose of 50 mg/kg. Treatments of cysteine-protected MoS2 dots almost restored BMNC was to
normal levels. *P<0.05 as compared with the control group (Student's t-test).
Figure S9 In vivo radiation protection of cysteine-protected MoS2 dots. (a) Hematological
analysis of irradiated mice treated with or without cysteine-protected MoS2 dots. Data were
collected at different time points of 1, 7 and 30 days after intraperitoneal injection (50 mg/kg,
n=10 mice/group). The results included red blood cells (RBC), white blood cells (WBC),
platelets (PLT), mean corpuscular hemoglobin (MCH), mean corpuscular hemoglobin
concentration (MCHC), mean corpuscular volume (MCV), hemoglobin (HGB), and hematocrit
(HCT). Data were analyzed using Student's t-test and * in (a) and (b) indicates p < 0.05.
Figure S10 In vivo radiation protection of cysteine-protected MoS2 dots. (b) Blood
biochemistry analysis of irradiated mice treated with or without cysteine-protected MoS2 dots at
different time points of 1, 7 and 30 days (50 mg/kg, n=10 mice/group). The results showed mean
and standard deviation of alanine aminotransferase (ALT), aspartate aminotransferase (AST),
total protein (TP), albumin (ALB), blood urea nitrogen (BUN), creatinine (CREA), globulin
(GOLB) and total bilirubin (TBIL). The results from hematology and biochemistry panels
showed the cysteine-protected MoS2 dots can recover the indicators 30 days post injection. Data
were analyzed using Student's t-test and * in (a) and (b) indicates p < 0.05.
Figure S11 Time-course renal clearance of cysteine-protected MoS2 dots. Time-dependent
concentrations of cysteine-protected MoS2 dots (50 mg/kg) in urine of mice at different time
points post injection, measured by ICP-MS.
Figure S12 In vivo toxicities of cysteine-protected MoS2 dots. (a) Hematological data of mice
treated with cysteine protected-MoS2 dots. Data were collected at different time points of 1, 7
and 30 days after intraperitoneal injection (50 mg/kg, n=8 mice/group). The results included PLT,
MCH, MCHC, MCV, HGB, and HCT.(b) Blood biochemistry analysis on mice treated with
cysteine-protected MoS2 dots at different time points of 1, 7 and 30 days. The results showed
mean and standard deviation of TP, ALB, BUN, CREA, GOLB, and TBIL. Data were analyzed
using Student's t-test and * in (a) and (b) indicates p < 0.05.
Figure S13 Representative pathological analysis for bladder, lung, heart and testis of mice
treated with cysteine-protected MoS2 dots (50 mg/kg, n=8 mice/group) at different time points of
1, 7 and 30 days. Scale bars, 50 µm.
|
1311.7138 | 1 | 1311 | 2013-11-27T15:15:36 | On the ion-mediated interaction between protein and DNA | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.BM"
] | The mechanism allowing a protein to search of a target sequence on DNA is currently described as an intermittent process composed of 3D diffusion in bulk and 1D diffusion along the DNA molecule. Due to the relevant charge of protein and DNA, electrostatic interaction should play a crucial role during this search. In this paper, we explicitly derive the mean field theory allowing for a description of the protein-DNA electrostatics in solution. This approach leads to a unified model of the search process, where 1D and 3D diffusion appear as a natural consequence of the diffusion on an extended interaction energy profile. | physics.bio-ph | physics | July 10, 2021
2:37
WSPC - Proceedings Trim Size: 9in x 6in
contribbarbi-v4
1
On the ion-mediated interaction between protein and DNA
M. Barbi
CNRS LPTMC UMR 7600, Universit´e Pierre et Marie Curie-Paris 6
4 place Jussieu, 75252 Paris Cedex 05, France
E-mail: [email protected]
Department of Chemistry, University of Cambridge, Lensfield Road
F.Paillusson
CB2 1EW, Cambridge, UK
E-mail: [email protected]
The mechanism allowing a protein to search of a target sequence on DNA
is currently described as an intermittent process composed of 3D diffusion in
bulk and 1D diffusion along the DNA molecule. Due to the relevant charge of
protein and DNA, electrostatic interaction should play a crucial role during
this search. In this paper, we explicitly derive the mean field theory allowing
for a description of the protein-DNA electrostatics in solution. This approach
leads to an unified model of the search process, where 1D and 3D diffusion
appear as a natural consequence of the diffusion on an extended interaction
energy profile.
Keywords: DNA; Proteins; ionic liquids; modelling
1. Introduction
Many proteins in living cells have to search specific, short sequences on
long DNA molecules in order to perform their biological task. Such DNA-
binding proteins have proven to be very efficient in searching their target:
their association constants can be two orders of magnitude higher than what
is expected from a simple 3D diffusion.1,2 It has been suggested3,4 that such
a rapid reaction rate can results from an intermittent diffusion, swapping
between a 1D diffusion along DNA -- or sliding -- and a 3D diffusion in
solution -- or jumping. An increasing number of single particle experiments
has been able to evidence sliding, confirming this scenario.5 -- 8 Experiments
also show that both the sliding and the jumping result to be sensitive to the
salt concentration.1,6,8 This supports the idea that electrostatics is involved
3
1
0
2
v
o
N
7
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
8
3
1
7
.
1
1
3
1
:
v
i
X
r
a
July 10, 2021
2:37
WSPC - Proceedings Trim Size: 9in x 6in
contribbarbi-v4
2
to some extent in the intermittent behaviour, with a probable role for the
solution ions.
In this paper, we shall first recall the statistical mechanics of ions in
solution for a given fixed charge distribution. We will then introduce a
toy model for a DNA-protein system9,10 for which we will discuss some
unexpected features. This will finally allow us to gain more insights about
the physics at play during the search of target by a protein and to propose
a method to get further insights on protein physical properties.
2. Statistical mechanics of electrolytes
In this section we consider two fixed macro-ions confined in a domain
Σ ⊂ R3 that contains an electrolyte solution. We denote ρf the charge
density carried by these macro-ions. Within the framework of statistical
mechanics,11 the position Rj of an ion j is a random vector that can take
any value ri belonging to a subset Ω of Σ. The set of N particle positions
{r1, .., rN} gotten at each trial for N ions in the system will be called an
ionic configuration and denoted C. In thermodynamic equilibrium, the con-
ditional probability density to get a specific configuration C knowing that
the system is at temperature β−1/kB and confined in the domain Ω of
volume V reads:
p(Cβ, N, V ) ≡ 1C∈ΩN e−βH(C)/Q[β,N,V ]
(1)
where 1C∈ΩN is the characteristic function that is zero if any of the ions is
outside Ω and one otherwise. In Eq. (1), we introduced the normalization
functional:
Q[β,N,V ] ≡
1(cid:81)m
(cid:90)
d3N (C) e−βH(C)
α
ΩN
α Nα!Λ3Nα
(2)
with d3N (C) being the 3N-dimensional Lebesgue measure on R3N , m is
the number of different ionic species in solution, Λα is the de Brooglie
wavelength of the species α and the sets N = {Nα}α=1..m ∈ Nm and
q = {qα}α=1..m ∈ Rm characterize the ionic composition of the mixture.
The real valued function H(C) in Eqs. (1) and (2) is the energy of the
system for a given configuration C that defines the model used. Here, we
rely on the so called Restricted Primitive Model (RPM),12 defined as:
(cid:88)
delta "function" to define the charge density ρC(r) ≡(cid:80)
d6rr(cid:48) ρC(r)ρC(r(cid:48))
4πεr − r(cid:48) +
H(C) =
where ε is the dielectric permittivity of water and where we use the Dirac
j eqjδ(r−rj)+ρf (r).
vHS(j, k)
(3)
(cid:90)
1
2
Σ
j<k
July 10, 2021
2:37
WSPC - Proceedings Trim Size: 9in x 6in
contribbarbi-v4
3
The second term in Eq. (2) stands for a hard-sphere repulsion such that
e−βvHS (j,k) behaves as Θ(rj − rk − D) where Θ(x) is the Heavyside step
function and D the diameter of the ionsa. From Eq. (2), one can then
perform a Hubbard-Stratonovich transform that effectively breaks Coulomb
pair interactions into a one-body potential φ12,13 b:
−1
r,r(cid:48) φ(r(cid:48))−iβ(cid:82) d3rφ(r)ρC(r)
(cid:82)
Σ d6rr(cid:48) φ(r)G
(cid:90) D[φ]
Σ d6rr(cid:48) ρC (r)ρC (r(cid:48))
− β
− β
(cid:82)
2
2
4πεr−r(cid:48) ≡
e
Z[0]
e
(4)
where i2 = −1 and G−1
r,r(cid:48) = −ε∆δ(r − r(cid:48)).13 The factor Z[0] is the nor-
malization factor for the free field φ in absence of ρC. If we introduce the
of Eq. (4) reads (cid:104)e−iβ(cid:82) d3r ρCφ(cid:105)φ. It is now convenient to swap to a grand
Gaussian average over configurations of the field φ, (cid:104).(cid:105)φ, then the r.h.s.
any value in Nm with a probability weight eβM·N ≡ eβ(cid:80)
canonical ensemble where the composition N is a random vector taking
α µαNα set by m
chemical potentials M = {µα}α=1..m ∈ Rm, each of which corresponds to
a particular ionic species. In this case, the normalization functional writes:
Q[β,M,V ] ≡(cid:88)
eβM·NQ[β,N,V ]
(5)
Inserting the partition function (2) (after performing (4)) into Eq. (5) yields:
(cid:68)
N
[β,M(iφ),V ]e−iβ(cid:82) d3r ρf φ(cid:69)
QHS
φ
Q[β,M,V ] =
(6)
[β,M(iφ),V ]
where QHS
is the grand partition function11 of a mixture of
bare hard spheres with the set of chemical potentials M(iφ) = M −
ieqφ. We now use the fact that the mixture is dilute by approximating
[β,M(iφ),V ] ≈
QHS
[β,M(iφ),V ] by the first term of its Mayer expansion11 i.e. QHS
α). Eq. (6) becomes then explicitely:
(cid:80)m
α=1 eβ(µα−ieqα φ)/Λ3
exp((cid:82) d3r (cid:80)
α eβ(µα−ieqαφ)/Λ3
e−β(cid:82)
(cid:90) D[φ]
2 −β−11r∈Ω
Σ d3r [ε (∇φ)2
α+iρf φ]
Q[β,M,V ] =
Z[0]
(7)
Note that the quadratic term in φ from Eq. (4) has now become quadratic
in ∇φ, by applying twice the divergence theoremc.
aThe diameter D is such that we can assume each ionic species to be in a stable gas
phase in solution.
bThe measure D[φ] can be thought of as the limit of the measure(cid:81)(L/+1)3
dγ(φk) --
γ being a complex measure -- characterizing field configurations on a 3D-lattice of size
L and lattice spacing when the latter tends to zero.
cThe boundary terms arising from this theorem do not contribute since global electro-
neutrality is assumed in Σ.
k
July 10, 2021
2:37
WSPC - Proceedings Trim Size: 9in x 6in
contribbarbi-v4
4
The so called Poisson-Boltzmann (PB) theory can now be readily
gotten from Eq. (7) by formally using a functional saddle point or
mean field approximation.12,14 The grand potential defined as G[β,M,V ] ≡
−β−1 ln Q[β,M,V ] reads then:
m(cid:88)
(∇ϕ)2
2
+ β−11r∈Ω
eβ(µα−h−eqαϕ)
α=1
Λ3
α
− ρf ϕ(cid:3) (8)
(cid:90)
Σ
d3r(cid:2)ε
G[β,M,V ]
M F= −
where the sign M F= stands for an equality within the mean field approxima-
tion and where ϕ ≡ iφs, φs being the field evaluated at the saddle point.
With the new field h introduced in (8) and ρf , G[β,M,V ] can be seen as a gen-
erating functional from which one can get grand canonical averages of mean-
ingful quantities. In particular (cid:104)iφ(cid:105)β,M,V ≡ −(δGβ,M,V /δρf )h=0
M F= ϕ and
(cid:104)ραC(cid:105)β,M,V ≡ −eqα(δGβ,M,V /δ[β(µ − h)])h=0
M F= eqα1r∈Ωeβµα e−βeqαϕ/Λ3
α.
The average charge density (cid:104)ραC(cid:105)β,M,V has to be a real number which implies
then that ϕ has to be a real field. It is therefore common to use Eq. (8) as a
functional of the real valued function ϕ called the Poisson-Boltzmann func-
tional that has to be extremalized numerically to find the most probable
electrostatic field ϕ and its corresponding charge density.15 An equivalent
way to look at it is to realize that if ϕ maximizes the PB functional: the
functional derivative of the latter with respect to the former has therefore
to be zero. This gives rise to an Euler-Lagrange type of equation called the
Poisson-Boltzmann equation to be solved for ϕ:
(cid:33)
∆ϕ = − 1
ε
1r∈Ω
eqα
eβµα
Λ3
α
e−βeqαϕ + ρf
(9)
(cid:32)
m(cid:88)
α=1
which turns out to be the most common route used to determine the average
electrostatic potential ϕ.
3. Modelling protein-DNA interactions
Many DNA binding proteins happen to have a concave shape matching
that of DNA. This is believed to optimize the recognition at the target
site. In addition, these proteins need to be positively charged otherwise
they would be repelled by DNA's high negative charge. In a previous work,
we have suggested a toy model to study the relevance of the geometry in
DNA-protein non specific interactions.9,10 In this model, depicted in Fig.
1 (a), the model DNA (MDNA) is a uniformly charged cylinder and the
model protein (MP) is a cylinder of larger radius but with an indentation
of cylindrical shape that matches exactly the DNA shape, and is positively
July 10, 2021
2:37
WSPC - Proceedings Trim Size: 9in x 6in
contribbarbi-v4
5
charged at the interface. They are also immersed in a RPM of a symmetric
1:1 electrolyte as described in Eq. (3). In the terminology of section (2),
the charge density on these macromolecules -- for a given distance L be-
tween them -- corresponds to ρf (L), the whole system is in a domain Σ
and Ω(L) is the accessible region to ions i.e. anywhere in Σ except inside
the macromolecules. The + and − ion bulk concentrations are assumed
to be the same and denoted nb ≡ eβµ± /Λ3±. The grand potential G[β,M,V ]
(a)
(b)
(c)
Fig. 1.
(a) Toy model for a protein interacting with a DNA segment in solution. (b)
The toy model can be mapped onto two oppositely but non-symmetrically charged plates
in solution. (c) Free energy profiles obtained by integrating the PB equation for the plate
plate system with (physiological) monovalent salt concentration of 0.1 mol/L, and for 4
different protein charge densities in the range from 0.06 to 0.3 times the absolute value
of the DNA charge density.
of this whole system is implicitly dependent on the domain Ω(L) and the
distribution ρf (L). In fact, if we were to consider L as a random variable
subject to thermal fluctuations, the grand potential G[β,M,V ](L) would act
as an effective interaction energy 16 between the MDNA and the MP such
that each value l of L appears with a probability weight e−βG[β,M,V ](l).
Monte Carlo (MC) simulations of the system described above have been
performed to compute exactly G[β,M,V ](L) via a so called thermodynamic
integration.17 It was found9,10 that G[β,M,V ](l) is an increasing function if
l ∈ [l∗, +∞[ and a decreasing function if l ∈ [0, l∗[. In physical terms,
July 10, 2021
2:37
WSPC - Proceedings Trim Size: 9in x 6in
contribbarbi-v4
6
l∗ corresponds to a stable equilibrium distance between the MP and the
MDNA segment.18 It was also shown that the profile G[β,M,V ](l) gotten
from MC simulations could be matched with a PB theory for two plates (as
shown in Fig. 1 (b)) by solving the PB equation (9) for ϕ and evaluating the
expression (8) for every l. This "mapping" between the MC implementation
of the toy model of Fig. 1 (a) and a PB treatment of a two plate system is
valid provided effective charge densities -- related to that of the MDNA and
the MP -- are used for the plates.9,10 The actual values of these parameters
depend on the particular modelling of ρf used in the MC simulations and
therefore do not provide at the moment any more insights about what is
happening in the system.
Overall, the intermittent behaviour observed for the DNA-protein sys-
tem can be rationalized by considering the random nature of L and treating
properly the physics of the ions: in a non-specific DNA-protein bound state,
sliding is possible at the equilibrium distance l∗, while thermal fluctuations
can still make the protein escape from DNA, in which case it would per-
form a jump. Let us finally note that, in practice, the density ρf cannot be
measured experimentally, and is often inferred from structural data. Inter-
estingly, the simple planar PB description that we have introduced is not
only able to capture this physics, but also provides analytical expressions --
as a function of effective charge densities ρf -- for both l∗ and G[β,M,V ](l∗).
Since these quantities directly determine the kinetic behaviour of the pro-
tein, a comparison with independent structural and dynamical data from
experiments may be used as a simple alternative to estimate coarse grained
surface densities at the protein-DNA interface.
References
1. A. D. Riggs, S. Bourgeois and M. Cohn, J. Mol. Biol. 53, 401 (1970).
2. P. H. Richter and M. Eigen, Biophysical Chemistry 2, 255 (1974).
3. P. von Hippel and O. Berg, J. Biol. Chem. 264, 675 (1989).
4. M. Coppey, O. B´enichou, R. Voituriez and M. Moreau, Biophys. J. 87, 1640
(2004).
5. N. Shimamoto, J. Biol. Chem. 274, 15293 (1999).
6. P. Blainey, A. van Oijen, A. Banerjee, G. Verdine and X. Xie, Proc. Natl.
Acad. Sci. U.S.A. 103, 5752 (2006).
7. J. Elf, G. Li and X. Xie, Science 316, 1191 (2007).
8. I. Bonnet, A. Biebricher, P.-L. Port´e, C. Loverdo, O. B´enichou, R. Voituriez,
C. Escud, W. Wende, A. Pingoud and P. Desbiolles, Nucleic Acids Research
36, 4118 (2008).
9. V. Dahirel, F. Paillusson, M. Jardat, M. Barbi and J.-M. Victor, Phys. Rev.
Lett. 102, p. 228101 (2009).
July 10, 2021
2:37
WSPC - Proceedings Trim Size: 9in x 6in
contribbarbi-v4
7
10. F. Paillusson, V. Dahirel, M. Jardat, J.-M. Victor and M. Barbi, Phys. Chem.
Chem. Phys. 13, 12603 (2011).
11. D. McQuarrie, Statistical mechanics (second revised edition) (University sci-
ence books, 2000).
12. J.-M. Caillol, J.Stat.Phys. 115, p. 1461 (2004).
13. D. Brydges and P. Martin, J.Stat.Phys. 96, p. 1163 (1999).
14. R. Netz and H. Orland, Eur. Phys.J. E 1, p. 203 (1999).
15. A. Maggs, Eur. Phys.Lett. 98, p. 16012 (2012).
16. G. Voth, Coarse-Graining of Condensed Phase and Biomolecular Systems
(CRC press, 2008).
17. D. Frenkel and S. B., Understanding molecular simulation (second edition)
(Academic Press, 2002).
18. F. Paillusson, M. Barbi and J.-M. Victor, Mol. Phys. 107, 1379 (2009).
|
1508.03512 | 1 | 1508 | 2015-08-14T14:26:59 | Magnetic fields facilitate DNA-mediated charge transport | [
"physics.bio-ph",
"q-bio.BM"
] | Exaggerate radical-induced DNA damage under magnetic fields is of great concerns to medical biosafety and to bio-molecular device based upon DNA electronic conductivity. In this report, the effect of applying an external magnetic field (MF) on DNA-mediated charge transport (CT) was investigated by studying guanine oxidation by a kinetics trap (8CPG) via photoirradiation of anthraquinone (AQ) in the presence of an external MF. Positive enhancement in CT efficiencies was observed in both the proximal and distal 8CPG after applying a static MF of 300 mT. MF assisted CT has shown sensitivities to magnetic field strength, duplex structures, and the integrity of base pair stacking. MF effects on spin evolution of charge injection upon AQ irradiation and alignment of base pairs to CT-active conformation during radical propagation were proposed to be the two major factors that MF attributed to facilitate DNA-mediated CT. Herein, our results suggested that the electronic conductivity of duplex DNA can be enhanced by applying an external MF. MF effects on DNA-mediated CT may offer a new avenue for designing DNA-based electronic device, and unraveled MF effects on redox and radical relevant biological processes. | physics.bio-ph | physics | Source: Biochemistry, Vol. 54, No. 21, pp. 3392-3399, 2015; DOI: 10.1021/acs.biochem.5b00295
Magnetic Fields Facilitate DNA-Mediated Charge Transport.
Jiun Ru Wong, Kee Jin Lee, Jian‐Jun Shu, and Fangwei Shao*.
Division of Chemistry and Biological Chemistry, School of Physical and Mathematical Sciences, Nanyang Techno‐
logical University, 21 Nanyang Link, Singapore 637371 (Singapore)
KEYWORDS : Charge transport, magnetic field effects.
Exaggerate radical‐induced DNA damage under magnetic fields is of great concerns to medical biosafety and to bio‐
molecular device based upon DNA electronic conductivity. In this report, the effect of applying an external magnetic field
(MF) on DNA‐mediated charge transport (CT) was investigated by studying guanine oxidation by a kinetics trap (8CPG)
via photoirradiation of anthraquinone (AQ) in the presence of an external MF. Positive enhancement in CT efficiencies
was observed in both the proximal and distal 8CPG after applying a static MF of 300 mT. MF assisted CT has shown sensi‐
tivities to magnetic field strength, duplex structures, and the integrity of base pair stacking. MF effects on spin evolution
of charge injection upon AQ irradiation and alignment of base pairs to CT‐active conformation during radical propaga‐
tion were proposed to be the two major factors that MF attributed to facilitate DNA‐mediated CT. Herein, our results
suggested that the electronic conductivity of duplex DNA can be enhanced by applying an external MF. MF effects on
DNA‐mediated CT may offer a new avenue for designing DNA‐based electronic device, and unraveled MF effects on redox
and radical relevant biological processes.
Introduction
Electronic coupling of the highly organized array of ar‐
omatic bases down the double helical DNA makes DNA a
promising biomaterial for the conduction of electrical
charges, a process termed as DNA‐mediated charge
transport (CT).1 Efficient DNA CT over at least 200 Å of a
well‐stacked duplex were readily observed.2 Variety of
approaches, from time‐resolved spectroscopic measure‐
ments,3 biochemical assays2a,2c,4 to electrochemical meth‐
ods5 indicated that the integrity of the base pair stacking
is a crucial factor in modulating CT process. Dynamic
conformation of a stack of 4~5 base pairs was proposed to
be the gating factor for electron hopping between adja‐
cent base pair domains.6 Both biological significance and
technological ramifications lie in the high sensitivity to
base pair integrity. DNA damage/repair and signaling
proteins may harvest DNA CT as a redox‐based method in
vivo for fast allocation to close vicinity of genomic
anomalies,7 while molecular electronics or biosensing
nanoapparatus analyze mutagenesis5b,8 and protein‐
nucleic acids interactions9 via electrochemical observa‐
tion of DNA CT.
The possible negative effect of magnetic field (MF) on
the genomic stability had long raised health concerns.
With rapid development and wide application of magnet‐
ic based medical instruments for diagnosis and therapeu‐
tics, these concerns were further extended to the medical
practice and healthcare. Despite having some progress
made in theoretical and experimental studies, efforts to
provide conclusive evidences to link the effect of external
MF to the biological system remained a contentious issue,
as there was a lack of consistent pattern in the MF expo‐
sure induced changes or damages to the cellular DNA.10
While the underlying mechanism responsible for the bio‐
logical system exerted by the external MF remains con‐
troversial, several in vitro magnetic field effects (MFE)
can be taken note of. Exposure to static MF alone induces
no detectable lethal effects on the cell viability and/or
DNA damage regardless of the MF strength.11 However,
MF along with other stimuli, such as biological oxidants,
reactive oxygen species (ROS) or ionizing radiation, could
cause enhanced oxidative damage to the cellular DNA.
The change in the responsiveness to MF in presence of
the oxidative stress led to the postulation that MF mani‐
fests, rather than induces oxidative consequences of ROS
radicals. This was not surprising as it was well document‐
ed that external MF has an extensive influence on the
course and kinetics of chemical reaction containing radi‐
cals (pairs).12 In addition, alignment of magnetically ani‐
sotropic biomacromolecules under high MF density has
also been reported to be an efficient approach to form
DNA films with define 3D orientations.13 Recent conduc‐
tive AFM experiments from Naaman’s group also reported
interesting manipulation of electron flow through mono‐
layer of biomacromolecules via permanent magnets,14
suggesting that MF may even affect immobilized duplex
DNA on electrode surface.
Though
it would provide essential
fundamental
knowledge to unravel the biological roles and expand
technological ramification of DNA‐mediated CT, the ef‐
fects of external MF on charge propagation through DNA
duplex and in turn how MFE affects CT‐promoted oxida‐
tive DNA damage have not been well understood. Herein,
we had designed a series of DNA systems to investigate
how external MF can affect the electronic properties of
nucleic acids and alter the yield of radical intermediates
formed during DNA‐mediated CT. Chemical decomposi‐
tion
8‐
cyclopropyldeoxyguanosine (8CPG) after the excitation of a
distant attached photooxidant, anthraquinone (AQ), is
used to determine the efficiency of CT process under the
absence and presence of an applied MF. Here, we ob‐
served that oxidative damages of 8CPG mediated by DNA
CT through well matched and mismatched DNA are sig‐
nificantly elevated.
kinetic
fast
hole
trap,
of
a
Results and Discussions
a)
b)
Figure 1. (a) (Top) DNA assemblies used in the present study.
(Bottom) Structure of anthraquinone derivative (Q) and hole
trap (8CPG) utilized. (b) Schematic illustration of DNA‐
mediated CT under the presence of an external MF. A MF
generated from two neodymium magnets was applied simul‐
taneously during the excitation of the photooxidant. An elec‐
tron hole is injected into the DNA duplex and is eventually
trapped by a hole trap to form oxidative damage.
Experimental Design. Anthraquinone (AQ) is select‐
ed as the photooxidant to trigger efficient charge
transport into the DNA. It had been well documented
that upon photoexcitation, the AQ moiety is able to un‐
dergo rapid intersystem crossing to generate long‐lived
triplet states that were capable of oxidizing the natural
nucleobases.15 The attachment of AQ to deoxyuridine via
an acetylene linker would ensure strong electronic cou‐
pling with the DNA (cid:0)‐stack and restrict the electronic
injection at the anchoring site. AQ1 and AQ2 contain two
guanine doublets, proximal and distal to the photooxi‐
dant, respectively. 5’‐G of the proximal GG doublet in AQ1
and that of the distal GG in AQ2 are replaced with 8CPG to
report the formation of guanine radical cation at corre‐
sponding GG sites via DNA CT (Figure 1a). Upon anneal‐
ing to either complementary DNA or RNA strands, DD
and DR, B‐form DNA or A‐form hybrid duplexes can be
obtained. Upon irradiation at 350 nm, the excited AQ
would be competent to abstract an electron from deoxy‐
uracil and inject an electron hole into the DNA (cid:0)‐stack.
The resulting radical cation would propagate along the
base pair stacks until it is trapped by 8CPG via a rapid ring
opening reaction to form permanent oxidative product
(Figure 1b). DNA‐CT yields can then be revealed by 8CPG
decomposition via HPLC analysis after the whole duplex
was enzymatically digested to nucleosides. DNA CT under
an external magnetic field (MF) was performed by placing
the samples between a pair of permanent neodymium
magnets, which provide a highly homogenous magnetic
flux across the aqueous samples (Figure 1b).
Figure 2. % Decomposition of 8CPG for B‐form DNA duplexes
(AQ1‐DD, AQ2‐DD) and A‐form DNA/RNA hybrid duplexes
(AQ1‐DR, AQ2‐DR) in 20 mM sodium phosphate buffer (pH
7.0) after irradiation for 10 min at 350 nm under the absence
(crossed) and presence (shaded) of an external MF.
Magnetic field effects accelerate 8CPG decomposi‐
tion via DNA CT. We first elucidate the influence of an
applied external MF on DNA‐mediated CT by monitoring
8CPG decomposition in DNA duplexes, AQ1‐DD and AQ2‐
DD (Figure 2). In the absence of an applied MF (B0
mT), 7 % and 12 % of 8CPG at the proximal (AQ1‐DD) and
distal (AQ2‐DD) G doublets were decomposed after 10
minutes’ irradiation, respectively. Under external MF
2
(B300 mT), both 8CPG decompositions were remarkably
elevated to 33% in AQ1‐DD and 44% in AQ2‐DD, respec‐
tively. Slightly higher increment was observed over longer
propagation length in AQ2‐DD (32%) than that in AQ1‐
DD (26%). Irradiation of either GG2‐DD, duplex without
photooxidant, or a mixture of AQ0‐DD and GG2‐DD, in
which AQ and 8CPG are placed in separate duplexes (Fig
S1), were conducted under both B0 mT and B300 mT.
No decomposition of 8CPG was observed above the noise
level in both control experiments. The fact that 8CPG re‐
mained intact in the absence of AQ, regardless of external
MF, indicated that 8CPG is decomposed only by photoin‐
duced oxidation from distant AQ, and MF does not cause
non‐redox damage to guanines. In the second control, AQ
in AQ0‐DD does not induce any 8CPG decomposition in
the separate duplex, GG2‐DD, even in the presence of
external MF. This suggested that no diffusible oxidants,
such as ROS, were causing damage to 8CPG under experi‐
mental conditions either with or without the presence of
MF. Thus the effects of external MF on stabilizing diffusi‐
ble radical oxidants and consequently enhance guanine
damage were invalid here. The elevation of 8CPG decom‐
position in AQ‐DD duplexes after the application of an
external MF was the sole consequence of MF effects on
DNA‐mediated charge transfer.
Figure 3. % Decomposition of 8CPG as a function of magnetic
flux density (B) for AQ1‐DD (×) and AQ2‐DD (ο) in 20 mM
sodium phosphate buffer (pH 7.0) after irradiation for 10 min
at 350 nm.
Dependence of 8CPG decomposition on magnetic
flux intensity. The dependence of 8CPG decomposition
on varying magnetic flux intensity (B0‐300 mT) upon
irradiation was further investigated and the results are
showed in figure 3. It appeared that the increment of
8CPG decomposition was dependent on the strength of the
magnetic flux density. Obvious MF effects on accelerating
DNA CT could be detected at B as low as 7 mT. CT yields,
in form of 8CPG decomposition, increased linearly initially
and approached plateau saturation after 100 mT for both
duplexes, AQ1‐DD and AQ2‐DD. Throughout the entire
flux intensity range, the elevation of 8CPG decomposition
efficiency at distal G doublets was always more pro‐
nounced than the proximal site. Better appreciation of
flux intensity increment over longer DNA bridge in AQ2‐
DD further confirmed that enhancement of guanine
damage as 8CPG decomposition is due to the effects of
external MF on oxidative DNA CT.
Structural dependence. It is well known that CT in
DNA exhibits different efficiencies in various secondary
structures of nucleic acids as measured by guanine oxida‐
tion,16 fluorescence quenching,17 electrochemistry18 and
transient absorption spectroscopy.19 Among the various
helical structures that DNA can adopt, A‐form DNA/RNA
hybrids are of great interest as they are essential to genet‐
ic transduction.20 Hence, we explored whether the mag‐
netic field effect was also applicable to A‐form DNA/RNA
hybrid duplex. AQ1‐DR and AQ2‐DR were formed by
annealing RNA version of complementary strands, DR, to
AQ1 and AQ2 (Figure 1a). CD spectra of AQ1‐DR and
AQ2‐DR showed a characteristic A‐form structure (Fig
S2). Upon irradiation, AQ1‐DR and AQ2‐DR showed a
decomposition of 2 % and 3 %, respectively (Figure 2). A
lower 8CPG decomposition observed in the A‐form hybrid
duplex was not unexpected, since the appreciable inter‐
strand stacking, due to low twist and large positive tilt, in
A‐form helices may not be optimal to DNA CT, as com‐
pared to intrastacking in B‐form duplex.16a,21 Albeit the
lower decomposition of 8CPG, under an external MF (B0
mT), AQ1‐DR and AQ2‐DR showed decent increments in
the damage of 8CPG to 6 % and 7%, respectively. Elevation
of CT yields was not as significant as those in B‐form du‐
plex, which is probably due to that hybrid helix with wid‐
er grooves and more compact structure may not be flexi‐
ble and dynamic enough to be tuned towards CT‐optimal
conformation by external MF. Furthermore, the higher
melting temperature value for the hybrid (AQ1‐DR and
AQ2‐DR) than the canonical B‐form duplexes (AQ1‐DD
and AQ2‐DD) indicated that the lower yield obtained was
not associated with a weaker structural stability (Table
S1). Regardless, MFE can also be observed in A‐form hy‐
brid duplexes and was not restricted specifically to B‐form
helix. In view of these results, the difference in the dam‐
age enhancement for both the proximal and distal GG
under the absence and presence of an external MF sug‐
gested that charge migration was aided by exposing du‐
plexes to an external MF.
Scheme 1. Simplified scheme showing the pathways
for photoinduced DNA CT in present work.
3
The origins of MFE on DNA CT. The complicated na‐
ture of a photoinduced DNA CT makes it challenging to
identify the reaction steps that were affected by the ex‐
ternal MF. Scheme 1 showed a simplified energy level dia‐
gram of three major stages of DNA CT in current system.
Charge injection (CI, stage 1) was initialized upon pho‐
toirradiation of AQ to the singlet excited state, 1AQ*,
which would rapidly undergo intersystem crossing (ISC)
to yield the excited long‐lived triplet state, 3AQ*. Subse‐
quent charge separation (CS) between 3AQ* and electron‐
ically conjugated deoxyuracil would generate the initial
triplet radical pair (3[AQ•‐–dU•+], 3RP) and a charge as a
radical cation was then injected into the base pair stack.
In the second stage, the radical cation migrated through
bridge base pairs and would eventually oxidize 8CPG to
form [8CPG•+]. In the final stage, 8CPG•+, as a radical cation
would undergo rapid ring opening reaction to trap the
charge and complete charge transport (Scheme 1). Since it
was well documented that MF could influence the spin
dynamics of radical intermediates in biological reac‐
tions,12 we postulated that external MF might affect
charge injection via 3RP and sequential migration of radi‐
cal cation, while the final stage was unlikely altered signif‐
icantly by external MF.22
Figure 4. % Decomposition of 8CPG as a function of time in s
for trinucleotides, 5’‐Q8CPGT‐3’, in 20 mM sodium phosphate
buffer (pH 7.0) after irradiation at 350 nm under the absence
(ο) and presence (×) of an external MF (300 mT).
In the charge injection stage, the propagating radical
cation on the DNA bridge, 3RP, (3[AQ•‐–dU•+]), may alter‐
natively cross over to singlet radical pair (1[AQ•‐–dU•+],
1RP) via triplet‐singlet (T–S) radical pair intersystem
crossing (RP–ISC). 1RP would then decay to the singlet
ground state by spin selective charge recombination (CR).
A trinucleotide, 5’‐Q8CPGT‐3’ (Figure 1a for structure) was
therefore used to elucidate whether MF can facilitate
charge injection by redistributing 3RP and 1RP popula‐
tions, via RP‐ISC, in AQ‐≡‐dU system. The trinucleotide is
an ideal assembly for such investigation. The flexible
phosphate chain would keep photooxidant, AQ, and hole
trap, 8CPG, in a close distance to ensure rapid charge injec‐
tion without being limited by diffusion. The electronic
coupling between AQ and 8CPG is strong enough to ensure
that RP remain as a germinate pair, and dissociation of
the radicals is suppressed to allow sufficient time for T–S
interconversion to develop under an applied MF. Further
charge migration is eliminated due to the lack of duplex
formation in short trinucleotide. Figure 4 showed the
decomposition of 8CPG in the trinucleotide in the absence
and presence of a 300 mT magnetic flux following 0‐60 s
of photoirradiation. Under background magnetic field
(B0 mT), up to 17 % of 8CPG undergo irreversible oxida‐
tive ring‐opening reaction with increasing irradiation to
60 s. Whereas under B300 mT, the trinucleotide con‐
sistently exhibit a higher decomposition (up to 23 %) dur‐
ing the entire irradiation time course. While our experi‐
mental setup may not be able make a distinction between
the singlet and triplet radical pairs, any variation in the
quantum yield of 3RP would directly affect the oxidative
decomposition of 8CPG due to the fact that 3RP has a much
longer lifetime than 1RP and should be the major species
for charge injection. The appreciable enhancement in
8CPG decomposition in trinucleotide indicated that the
stabilization of spin‐correlated RP, specifically the triplet‐
state, is one of MF effects on DNA CT. Such MFE can be
interpreted with reference to a simple but well‐known RP
model.12 Assuming that the exchange interaction energy
between the 1RP and 3RP is very small, at zero applied
field, the singlet (S) and the 3 triplet sublevels (T0, T±1) are
nearly isoenergetic and there will be unrestricted RP‐ISC
between the 4 states, induced by electron‐nuclear hyper‐
fine interaction (HFI). As the applied MF strength in‐
creased, the electronic Zeeman interaction would slow
down and uncoupled the T±1 from the conversion process.
Eventually, only the interconversion between T0–S states
would be remained at high flux intensity (hyperfine
mechanism). Hence the conversion of 3RP to 1RP was di‐
minished and decay via charge recombination to the sin‐
glet state would be impeded. Consequently, the popula‐
tions of 3RP would increase and charge injection from 3RP
as an ‘escape’ product from CR would be enhanced and be
revealed as more hole trapping at 8CPG site. Hence in tri‐
nucleotide model, higher efficiency of 8CPG decomposi‐
tion under external MF implied that charge injection is
facilitated by MF via enriching the triplet radical pair,
3[AQ•‐–dU•+] during photoinitiated charge separation.
Figure 5. % Decomposition of 8CPG for AQ1‐ and AQ2‐CA, CC
or CT in 20 mM sodium phosphate buffer (pH 7.0) after irra‐
4
diation for 10 min at 350 nm under the absence (crossed) and
presence (shaded) of an external MF.
Notably, it was well accepted that MF could induce ori‐
entation in organic molecules23 and biological macromol‐
ecules.24 The preferential orientation originates from the
external magnetic induction and the anisotropy suscepti‐
bility of the DNA nucleic bases.23,25 Although the magnet‐
ic susceptibility of neutral nucleobases are weak, para‐
magnetic radical cation as the propagating species in
DNA CT, could exhibit a much stronger response to MF.
An intervening mismatch is a severe disruption to the
integrity of DNA base pair stacking and could significant‐
ly diminish the overall CT yields.2a,8,26 Hence we challenge
the ability of MF to facilitate DNA CT by introducing dif‐
ferent mismatches to DNA bridge before and after 8CPG
trap. Upon annealing to complementary strands with sub‐
stitution of G by either A, C or T, duplexes containing CA,
CC and CT mismatches between proximal and distal GG
were formed and submitted to photoinitialized DNA CT
under B0 mT and B300 mT (Figure 1a). When AQ1‐
CA, AQ1‐CC and AQ1‐CT were excited under B0 mT,
8CPG at the proximal GG site in all three duplex showed a
decomposition of ~7 %. After B300 mT was applied,
8CPG damage was increased to ~30 % (Figure 5). These
results were almost identical to AQ1‐DD. None of the
mismatches diminished DNA CT between photooxidant
and proximal GG, which was not unexpected as the mis‐
matched base pair was positioned after the hole trap and
the intervening base pair stack between AQ and 8CPG was
not disturbed. With the photoinduced DNA CT remain‐
ing intact and similar degree of efficiency enhancement
being observed, it implied that MFE was not restricted by
the presence of a mismatched site in the duplex. Where‐
as, in the cases of AQ2‐CA, AQ2‐CC and AQ2‐CT, damage
yield of distal 8CPG under B0 mT was almost fully
quenched by a single base interruption. Similar to previ‐
ous studies on CT chemistry,8a,8b,26 base replacement of G
with A, C or T disrupted the integrity of the (cid:0)‐stack, and
inhibited charge propagation to the distal 8CPG. This
showed that CT under our DNA assembles is still sensi‐
tive to the integrity of base pair stacking as previously
reported. Interestingly, under B300 mT, decomposition
yields of 8CPG were recovered in AQ2‐CA (4 %), AQ2‐CC
(11 %) and AQ2‐CT (11 %). In the cases of two pyrimi‐
dine/pyrimidine mismatches, CT yield was restored to a
similar efficiency as the well‐matched DNA under back‐
ground MF. The reinstallation of DNA CT by external MF
suggested that MF had applied a well‐pronounced com‐
pensation effect to repair or shield the distortion of base
pair stacking. Characterization studies by X‐ray crystal‐
lography and NMR methods had shown that mismatch
base pair cause minimum alterations on the global con‐
formation of the B‐DNA duplex, but the distortions were
localized in the vicinity of the mismatched site.27 Hence,
it is likely that a MF as weak as 300 mT might be suffi‐
cient to confer partial base orientation to the mismatched
and the neighboring base pairs so that the optimal (cid:0)‐
stacking in the local environment can be recovered tem‐
porarily to the CT‐active conformations comparable to a
well‐matched duplex and sequentially efficient CT can be
assessed. Though the current experiments cannot distin‐
guish whether the compensation effects adjusted the dy‐
namic conformation of duplex when DNA was still in
neutral form or when DNA was oxidized radical and cati‐
on was transiently occupying the domain of base pairs, we
tends to believe that the latter should be more significant
due to low flux intensity we applied in the experiments.
The ability to accelerate CT in duplex DNA by the ap‐
plication of a weak MF could be the basis for designing a
magnetically controlled DNA electronic switch. The mag‐
nitude of the MF‐induced signal could be manipulated by
changing the nature of the mismatch base pair or the
DNA secondary structures. In a mismatched duplex, op‐
timum base stacking was not achieved and charge propa‐
gation would be shut off beyond the mismatch site.
Without applying MF, the mismatched duplex wire will
be in an “off” state. Likewise, switching on the MF with‐
out irradiation would not create any detectable DNA‐
mediated CT signal. However, when two triggering fuels,
an excitation source and an applied MF, were concurrent‐
ly present, the switch can then be turn “on”. Subsequent
removal of either fuel would restore the “off” state. Cou‐
pled with electrochemical device, MF alone could also be
used as a switch control for currents.
Given that DNA‐mediated CT can report on the integri‐
ty of DNA over long molecular distances, it had been re‐
cently proposed that DNA‐bound repair protein with a
metal cluster as the redox cofactor might exploit DNA CT
to signal and communicate with each other and so as to
efficiently detect lesions across the genome.7c Interesting‐
ly, our results suggested that with the assistance of MF,
local alignment near the mismatch site might permit the
charge migration to proceed without obstruction. As
such, electron may still be able to shuttle through the
damaged DNA between the two repair proteins in the
presence of a mild MF. Consequently, the proteins may
detach and bind to an alternative DNA site without first
locating and repairing the damage. In essence, the pro‐
teins will erroneously pass on false information about
DNA structural integrity to one another through the tran‐
siently aligned DNA (cid:0)‐stack by MF. Potentially, this could
be a reasonable justification why more DNA damages
were detected in the presence of an applied MF in the
biological systems. In fact, the number of lesion sites gen‐
erated by ROS could remain unchanged except that in the
presence of MF, the repair system was compromised since
repair enzymes were misled by MF‐assisted DNA CT and
bypass the screening of the lesion sites, leading to an “ac‐
cumulation” of damage spots.
Conclusion
Here by using anthraquinone as photooxidant and 8CPG
as hole trap, CT efficiencies through duplex DNA were
explored in the presence of an external MF. The applica‐
tion of MF caused an increased damage to 8CPG via photo‐
initialized DNA CT over various lengths of duplex bridge.
5
Such effects were also observed in the DNA/RNA hybrid
duplex, albeit lower yields. The acceleration of CT is
closely related to the strength of the applied MF, where
damage to 8CPG was detectable as low as 7 mT. Two fac‐
tors may account for the observed MF‐induced accelera‐
tion of DNA CT. Application of an external MF interfered
the spin evolution of RP and consequently, enhanced the
of the triplet state population and enhanced hole injec‐
tions into the DNA (cid:0)‐stack. Secondly, MF can manipulate
the alignments of base pair stacking to promote high ac‐
cessibility of DNA bridge to CT‐active conformation.
Herein, our results suggest that external MF could be a
promising approach to enhance the electronic conductivi‐
ty of duplex DNA. This approach would have open a new
venue to design DNA CT based molecular device and in‐
spire a better understanding of related biological and
medical processes.
Experimental Section.
Oligonucleotides synthesis. Cyanoethyl phospho‐
ramidite of 8‐cyclopropylguanosine28 were synthesized as
described while anthraquinone‐5‐ethynyl‐dU was pur‐
chased from Berry & associates. DNA oligonucleotides
with trityl‐on were synthesized using standard phospho‐
ramidite protocols on Bioautomation Mermade 4 DNA
synthesizer with reagents from Glen research. After incu‐
bation in AMA at 37 ºC for 2 h, the cleaved DNA strands
were purified by reverse phase HPLC (mircosorb 100‐5
C18 Dynamax, 250 × 10.0 mm), detritylated wth 80% gla‐
cial acetic acid for 15 min and repurified by reverse‐phase
HPLC. All the DNA oligonucleotides were confirmed by
ESI mass spectrometry and quantified by UV‐vis spectros‐
copy.
Photooxidation experiment. DNA duplexes (10 ,
30 L in 20 mM sodium phosphate buffer, pH 7.0) were
prepared by annealing the modified DNA strands with its
complements (ratio of 1: 1.1) and gradually cooled to room
temperature overnight after heating for 5 min at 90 ºC.
The duplexes were then irradiated with a 450 W Xenon
lamp, equipped with monochromater and a 320 nm long‐
pass filter for 10 min. Exposure to a stationary magnetic
field was performed by placing 2 neodymium magnets
(0.7 cm in diameter) on each side of the mircotube that
contained the samples. The opposite poles of the magnets
were separated by about 0.7 cm. As the surface area of the
magnetic disc was larger than that of the sample area, it
was considered that the magnetic field strength of ~ 300
mT between the opposite poles was homogenous. Follow‐
ing, the samples were digested into free nucleosides by
incubating at 37 ºC with phosphodiesterase I and alkaline
phosphatase for 24 h. The nucleosides were subsequently
separated and analyzed by reverse‐phase HPLC (Chem‐
cobond 5‐ODS‐H, 4.6 × 150 mm).
ASSOCIATED CONTENT
Supporting Information. Supporting figures and table as‐
sociated with this article is available free of charge via the
Internet at http://pubs.acs.org.”
AUTHOR INFORMATION
Corresponding Author
* E‐mail: [email protected]
ACKNOWLEDGMENT
The financial support for this research work by Ministry of
Education of Singapore (M4011040, M4020163) and Nanyang
Technological University (M4080531) is greatly appreciated.
ABBREVIATIONS
CT, charge transport; MF, magnetic field; MFE, magnetic
field effects; ROS, reactive oxygen species; 8CPG, 8‐
cyclopropyldeoxyguanosine; AQ, anthraquinone; CI; charge
injection; ISC, intersystem crossing; CS, charge separation;
RP, radical pair; CR, charge recombination.
REFERENCES
(1) (a) Delaney, S.; Barton, J. K. J. Org. Chem. 2003, 68, 6475;
(b) Giese, B. Annu. Rev. Biochem. 2002, 71, 51; (c) Lewis, F. D.;
Letsinger, R. L.; Wasielewski, M. R. Acc. Chem. Res. 2001, 34, 159;
(d) Schuster, G. B. Acc. Chem. Res. 2000, 33, 253.
(2) (a) Núñez, M. E.; Hall, D. B.; Barton, J. K. Chem. Biol. 1999,
6, 85; (b) Henderson, P. T.; Jones, D.; Hampikian, G.; Kan, Y. Z.;
Schuster, G. B. Proc. Natl. Acad. Sci. U.S.A. 1999, 96, 8353; (c)
Sanii, Ly, L.; Schuster, G. B. J. Am. Chem. Soc. 1999, 121, 9400.
(3) (a) Wan, C.; Fiebig,; T. Kelley, S. O.; Treadway, C. R.; Bar‐
ton, J. K.; Zewail, A. H. Proc. Natl. Acad. Sci. U.S.A 1999, 96, 6014;
(b) Takada, T.; Kawai, K.; Cai, X.; Sigimoto, A.; Fujitsuka, M.;
Majima, T. J. Am. Chem. Soc. 2004, 126, 1125; (c) Lewis, F. D.;
Daublain, H. Zhu, P.; Cohen, B.; Wasielewski, M. R. Angew.
Chem. Int. Ed. 2006, 45, 7982.
(4) (a) Hall, D. B.; Holmlin, R. E.; Barton, J. K. Nature 1996,
382, 731; (b) Ghosh, A.; Joy, A.; Schuster, G. B.; Douki, T.; Cadet,
J. Org. Biomol. Chem. 2008, 6, 916; (c) Nakatani, K.; Saito, I. Top.
Curr. Chem. 2004, 236, 163.
(5) (a) Gorodetsky, A. A.; Barton, J. K. Langmuir 2006, 22,
7917; (b) Kelley, S. O.; Hill, M. G.; Barton, J. K. Angew. Chem. Int.
Ed. 1999, 38, 941; (c) Drummond, T. G.; Hill, M. G.; Barton, J. K. J.
Am. Chem. Soc. 2004, 126, 15010.
(6) (a) O’Neill, M. A.; Barton, J. K. J. Am. Chem. Soc. 2004,
126, 11471; (b) Shao, F.; Augustyn, K.; Barton, J. K. J. Am. Chem.
Soc. 2005, 127, 17445.
(7) (a) Merino, E. J.; Boal, A. K.; Barton, J. K. Curr. Opin. Chem.
Biol. 2008, 12, 229; (b) Genereux, J. C.; Boal, A. K.; Barton, J. K. J.
Am. Chem. Soc. 2010, 132, 891; (c) Sontz, P. A.; Muren, N. B.; Bar‐
ton, J. K. Acc. Chem. Res. 2012, 45, 1792.
(8) (a) Boon, E. M.; Ceres, D. M.; Drummond, T. G.; Hill, M.
G.; Barton, J. K. Nat. Biotechnol. 2000, 18, 1096; (b) Kelley, S. O.;
Boon, E. M.; Barton, J. K.; Jackson, N. M.; Hill, M. G. Nucleic
Acids Res. 1999, 27, 4830; (c) Drummond, T. G.; Hill, M. G.; Bar‐
ton, J. K. Nat. Biotechnol. 2003, 21, 1192.
(9) Boon, E. M.; Salas, J. E.; Barton, J. K. Nat. Biotechnol. 2002,
20, 282.
(10) (a) Miyakoshi, J. Prog. Biophys. Mol. Biol. 2005, 87, 213; (b)
Miyakoshi, J. Sci. Technol. Adv. Mater. 2006, 7, 305.
(11) (a) Zmyslony, M.; Palus, J.; Jajte, J.; Dziubaltowska, E.; Raj‐
kowska, E. Mutat. Res. 2000, 453, 89; (b) Jajte, J.; Grzegorczyk, J.;
Zmyslony, M.; Rajkowska, E. Bioelectrochemistry 2002, 57, 107.
(12) (a) Steiner, U.E.; Ulrich, T. Chem. Rev. 1989, 89, 51; (b)
Grissom, Ch. B. Chem. Rev. 1995, 95, 3; (c) McLauchlan, K. A.;
6
Steiner, U. E. Mol. Phys. 1991, 73, 241; (d) Brocklehurst, B. Chem.
Soc. Rev. 2002, 31, 301.
(13) Morii, N.; Kido, G.; Suzuki, H.; Nimori, S.; Morii, H. Biom‐
acromolecules 2004, 5, 2297.
(14) (a) Göhler, B.; Hamelbeck, V.; Markus, T. Z.; Kettner, M.;
Hanne, G. F.; Vager, Z.; Naaman, R.; Zacharias, H. Science 2011,
331, 894; (b) Xie, Z.; Markus, T. Z.; Cohen, S. R.; Vager, Z.;
Gutierrez, R.; Naaman, R. Nano Lett. 2011, 11, 4652.
(15) Armitage, B.; Yu, C.; Devadoss, C.; Schuster, G. B. J. Am.
Chem. Soc. 1994, 116, 9847.
(16) (a) Odom, D. T.; Barton, J. K. Biochemistry 2001, 40, 8727;
(b) Sartor, V.; Henderson, P. T.; Schuster, G. B. J. Am. Chem. Soc.
1999, 121, 11027; (c) Núñez, M. E.; Noyes, K. T.; Gianolio, D. A.;
McLaughlin, L. W.; Barton, J. K. Biochemistry 2000, 39, 6190; (d)
Kan, Y.; Schuster, G. B. J. Am. Chem. Soc. 1999, 121, 11607.
(17) O’Neill, M. A.; Barton, J. K. J. Am. Chem. Soc. 2002, 124,
(18) Boon, E. M.; Barton, J. K. Bioconjugate Chem. 2003, 14,
13053.
1140.
(19) (a) Kawai, K.; Osakada, Y.; Sugimoto, A.; Fujitsuka, M.;
Majima, T. Chem. Eur. J. 2007, 13, 2386; (b) Choi, J.; Park, J.;
Tanaka, A.; Park, M. J.; Jang, Y. J.; Fujitsuka, M.; Kim, S. K.;
Majima, T. Angew. Chem. Int. Ed. 2013, 52, 1134.
(20) (a) Adams, R. L. P.; Knowler, J. T.; Leader, D. P. The Bio‐
chemistry of Nucleic Acids, 10th ed.; Chapman and Hall: London,
1986; Chapter 6; (b) Hansen, U. M.; McClure, W. R. J. Biol. Chem.
1980, 255, 9564; (c) Varmus, H. Science 1988, 240, 1427; (d) Ste‐
phenson, M. L.; Zamenick, P. C. Proc. Natl. Acad. Sci. U.S.A.
1978, 75, 285; (e) Varmus, H. E. Science 1982, 216, 812; (f) Ogawa,
T.; Okazaki, T. Annu. Rev. Biochem. 1980, 49, 421; (g) Kirkpatrick,
D. P.; Radding, C. M. Nucleic Acids Res. 1992, 20, 4347.
(21) (a) Saenger, W. Principles of Nucleic Acid Structure;
Springer‐Verlag: New York, 1984. (b) Hartman, B.; Lavery, R. Q.
Rev. Biophys. 1996, 29, 309.
(22) Unpublished results.
(23) Lohman, J. A. B.; Maclean, C. Chem. Phys. 1978, 35, 269;
Chem Phys 1979, 43, 144
(24) (a) Murayama, M. Nature 1965, 206, 420; (b) Mayer, A.;
Dransfeld, K. Phys. Rev. Lett. 1975, 35, 397; (c) Sugiyama, J.;
Chanzy, H.; Maret, G. Macromolecules 1992, 25, 4232; (d)
Worcester, D. L. Proc. Natl. Acad. Sci. U.S.A. 1978, 75, 5475.
(25) Veillard, A.; Pullman, B.; Berthier, G. C. R. Acad. Sci. 1961,
252, 2321.
(26) (a) Bhattacharya, P. K.; Barton, J. K. J. Am. Chem. Soc.
2003, 123, 8649; (b) Kelley, S. O., Holmlin, R. E.; Stemp, E. D. A.;
Barton, J. K. J. Am. Chem. Soc. 1997, 119, 9861.
(27) (a) Brown, T.; Kennard, O.; Kneale, G.; Rabinovich, D.
Nature 1985, 315, 604; (b) Hunter, W. N.; Brown, T.; Kneale, G.;
Anand, N. N.; Rabinovich, D.; Kennard, O. J. Biol. Chem. 1987,
262, 9962; (c) Hunter, W. N.; Brown, T.; Kennard, O. Nucleic
Acids Res. 1987, 15, 6589; (d) Sowers, L. C.; Fazakerley, G. V.;
Eritja, R.; Kaplan, B. E.; Goodman, M. F. Proc. Natl. Acad. Sci.
U.S.A, 1986, 83, 5434; (e) Fazakerley, G. V.; Quignard, E.;
Woisard, A.; Guschlbauer, W.; Vandermarbel, G. A.; Vanboom, J.
H.; Jones, M.; Radman, M. EMBO J. 1986, 5, 3697.
(28) Wong, J. R.; Shao, F. ChemBioChem 2014, 15, 1171.
Insert Table of Contents artwork here
7
|
0909.1740 | 2 | 0909 | 2010-03-19T16:55:08 | Redundancy and error resilience in Boolean Networks | [
"physics.bio-ph",
"cond-mat.stat-mech",
"nlin.CG"
] | We consider the effect of noise in sparse Boolean Networks with redundant functions. We show that they always exhibit a non-zero error level, and the dynamics undergoes a phase transition from non-ergodicity to ergodicity, as a function of noise, after which the system is no longer capable of preserving a memory if its initial state. We obtain upper-bounds on the critical value of noise for networks of different sparsity. | physics.bio-ph | physics | Redundancy and error resilience in Boolean Networks
Institut für Festkörperphysik, TU Darmstadt, Hochschulstrasse 6, 64289 Darmstadt, Germany
(Dated: November 2, 2018)
Tiago P. Peixoto∗
We consider the effect of noise in sparse Boolean Networks with redundant functions. We show that they
always exhibit a non-zero error level, and the dynamics undergoes a phase transition from non-ergodicity to
ergodicity, as a function of noise, after which the system is no longer capable of preserving a memory of its
initial state. We obtain upper-bounds on the critical value of noise for networks of different sparsity.
0
1
0
2
r
a
M
9
1
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
2
v
0
4
7
1
.
9
0
9
0
:
v
i
X
r
a
PACS numbers: 89.75.Da,05.65.+b,91.30.Dk,91.30.Px
Introduction -- Biological systems are unavoidably noisy in
their nature, but often need to function in a predictable fash-
ion [1]. In such a situation, strategies to diminish the harm-
ful effect of noise will significantly impact the fitness of a
given organism. The most fundamental protection mechanism
a system can adopt is the redundancy of its underlying com-
ponents, since the resulting coincidences necessary to impact
the proper function of the system can drastically diminish the
probability of error. In this letter we are concerned with the ef-
fect of redundancy in gene regulation; in particular in a simple
Boolean Network (BN) model. We assume that each compo-
nent in the system is arbitrarily redundant, with the only re-
striction that the number of inputs per component is fixed and
finite. In a general manner, we are able to show that redun-
dancy can always guarantee reliable dynamics, up to a given
critical value of noise, above which the system is incapable of
maintaining any memory of its past states. From simple con-
siderations, we are able to obtain upper bounds on the maxi-
mum resilience attainable. This provides an important frame
of reference to determine the reliability of a system with a
given sparsity.
We begin by defining the model, and how noise is intro-
duced. A Boolean Network (BN) [2] is a directed graph with
N nodes, representing the genes, which have an associated
Boolean state σi ∈ {0, 1}, corresponding to the transcription
state, and a function fi({σj}i), which determines the state of
node i given the states of its input nodes {σj}i. The number
of inputs of a given node is ki, or simply k if its the same
for all nodes. This system is usually updated in parallel, such
that at each time step t, we have σi(t + 1) = fi({σj (t)}i).
Starting from an initial configuration, the system will evolve,
and eventually settle on an attractor. In a real system, the ex-
pression level of a particular gene can fluctuate, despite the
stability of its input states [3]. This characteristic can be in-
corporated qualitatively in the BN model as uniform noise [4 --
10], defined as a probability p that, at each time step, the value
of a given input σj ∈ {σj}i of a node i is flipped, prior to the
evaluation of the function fi. The value of p plays the role of
a temperature in the system. If p = 0 the original determin-
istic model is recovered, and if p = 1/2 the system becomes
effectively decoupled, with entirely stochastic dynamics.
In the model above, it is known that error resilience does
not spontaneously emerge, since Random Boolean Networks
(BNs with random topology and functions [11]), and simple
functional elements such as loops always exhibit ergodic be-
haviour in the presence of noise (p > 0) [10]. To obtain re-
silience, some level of functional redundancy must be intro-
duced in the network. In the following we describe how this
can be done, and analyse the optimal situation where all func-
tions are arbitrarily redundant. From this situation we obtain
upper bounds on the maximum reliability attainable, which
is characterized by a transition from non-ergodic to ergodic
behaviour at a critical noise value.
Redundancy in sparse networks -- Given a finite number of
inputs per node k ≪ N , for any given (non-constant) func-
tion, the probability that noise will change the output of the
function will always be above zero, independent of the size
of the network. Therefore, in average there will always be a
non-vanishing fraction of the nodes which will be at the wrong
state at any given time. The most that can be expected is that
this fraction be as small as possible, and remain small as the
dynamics evolve. The issue of achieving the first goal was
first approached by von Neumann [12, 13], who described a
general mechanism of optimal redundancy, which is capable
of reducing the propagation of errors in a BN. We will briefly
outline this mechanism, and then show how it can be used to
construct a dynamical model of error propagation in resilient
Boolean Networks.
The mechanism proposed in [12] consists of locally repli-
cating a given function, such that the replicated input and out-
put edges will form bundles which will all carry the same in-
formation, in the absence of noise (see Fig 1). The edges of
f
f
f
f
f
r
a
n
d
o
m
r
r
r
r
Executive
Restoring organ
organ
f
FIG. 1: Redundancy construction method. Left: The original func-
tion in the network; Middle: Equivalent redundant function, com-
posed of the executive and restoring organs, with an edge bundle of
size four. The grey rectangle corresponds to a random rewiring of
the edges; Right: The resulting "pseudo-function", with incoming
and outgoing edge bundles.
the output bundle are then randomized and fed into appro-
priate restoration functions which will independently query
the majority state carried in them. The number of edges in
the bundle can be arbitrarily large, but the number of inputs
per node must remain fixed. The output bundle of the last
functions will then propagate the information to the rest of
the network, which is also modified in the same manner. The
first stage was dubbed in [12] the "executive organ", and the
second stage, the "restoring organ". We note that while this
method outlines a specific construction, it has a general na-
ture, since it incorporates the two the most necessary features
to be resilient against noise: replication and restoration of ma-
jority values. It does so piecewise for all functions in a given
network, depicting an alternative version with an optimal level
of redundancy. This robust version will then function exactly
like the original network, if each executive and restoring or-
gan is thought of as an individual function (which we will
call a pseudo-function). However, in the presence of noise,
the fraction b of outputs having value 1 (and conversely 1 − b
with value 0) that enter and leave such a pseudo-function is
no longer a Boolean value, but instead are real values in the
range [0, 1], which will be continuously distributed in the limit
of large number of edge in the bundles. In the following, we
assume that the number of edges in the bundle is large enough,
so that the fluctuations of the values of bi can be neglected.
In this case, these pseudo-functions will be generalizations of
the original Boolean function (plus restoration) in the real do-
main, which will regulate how noise is propagated in the net-
work. The general form of those pseudo-functions is
g({bi}) = r
2k−1
k−1
Xj=0
δf (j),1
Yi=0
(δji,1bi + δji,0(1 − bi))
,
(1)
where δij is the Kronecker delta, bi is the input bundle i, f is
the function of the executive organ, the variable j represents
a given Boolean input combination, ji the i-th bit of the in-
put combination j, and r(b) performs the restoring function
(which we will describe in detail below).
We are interested in analysing how the above mechanism
can hinder error propagation on the network. The effect of
noise can be measured in a variety of ways, but here we are
interested in the ability of the system in remembering its past
states, which we will label as non-ergodicity. More precisely,
we can define an order parameter, the long-term hamming dis-
tance,
h = lim
T →∞
1
T
T
Xt=0(cid:10)(cid:12)(cid:12)bi(tσa
i ) − bi(tσb
i )(cid:12)(cid:12)(cid:11) ,
(2)
where bi(tσ) is the value of b for pseudo-node i at time t,
with a starting state bi(0σ) = σ. The average h. . .i is taken
over the whole network, and several independent realizations
i}. If
of the dynamics, with different initial states {σa
the system shows ergodic dynamics (as previously discussed),
the value of h should converge to zero, corresponding to only
i } and {σb
2
one possible fixed point in the values of bi. Otherwise, it
should decrease with the noise strength p, as the effects of
noise brings the system closer to the ergodic phase [17].
We will consider separately the case of networks with func-
tions k = 2, and later the case k > 2. The case k = 1 will
not be analysed since it does not allow for the construction of
a restoring organ.
k = 2 -- The most crucial part in the procedure outlined
above is the selection of the function to be used in the restor-
ing organ. The function must be able to transform the values
of the majority of the edges in the input bundle, into a even
greater majority in the output bundle. However, no k = 2
function can act as a simple majority function. There are how-
ever some functions which behave as a majority function for
certain input combinations, but not for others. These functions
are the AND (8), NAND (7), OR (14) and NOR (1) which
react only to two simultaneous input changes, if the original
inputs are all 1 or 0, but react to any input flip if the origi-
nal inputs are in the opposite state. Therefore those functions
would be able to correct either value passing on the bundle,
but not both. The solution proposed in [12] is to construct the
restoring organ with two modules connected in sequence, both
with either NAND or NOR functions [18], as can be seen in
Fig. 2. The first tier will correct one of the values if it can
r
a
n
d
o
m
r
a
n
d
o
m
2
)
ǫ
◦
d
n
a
n
b
(
1.0
0.8
0.6
0.4
0.2
0.0
2
)
ǫ
◦
r
o
n
b
(
1.0
0.8
0.6
0.4
0.2
0.0
p = 0.01
p = 0.08
p = 0.2
r
a
n
d
o
m
r
a
n
d
o
m
p = 0.01
p = 0.08
p = 0.2
0.0
0.2
0.4
0.6
0.8
1.0
0.0
0.2
0.4
0.6
0.8
1.0
η
η
FIG. 2: Restoration organs with k = 2, with NAND (left) and NOR
(right) functions. Below each are the restoration maps, for some val-
ues of p. The dashed lines correspond to b(η) = η.
and intrinsically flip the majority value of the bundle, and the
second tier will then have its chance of correcting, now that
the majority value has changed. We can verify the actual re-
sponse of this scheme to noise, by defining two maps. First,
the actual noise on the bundle,
ǫ(η) = (1 − 2p)η + p,
(3)
where ǫ is the fraction of the edges in the bundle with a given
value, given the original fraction of same value η. Second, the
response of the NOR and NAND functions,
bnor(η) = (1 − η)2, bnand(η) = 1 − η2,
(4)
where b is the fraction of edges in the output bundle with value
1, and η the fraction of inputs with the same value. The full
restoration map is then given by
br(η) = (bnor/nand ◦ ǫ)2(η),
(5)
which is plotted in Fig. 2, for some values of p. We can see
that the majority value on the bundle is preserved, even for
non-zero values of p. However, the question remains if this
restoration will be enough to maintain trajectories of a net-
work from diverging. For that, we need to couple the restora-
tion map above with the functions present on the network and
iterate the system. However, there is one specific situation
which represents the limiting case of maximum resilience,
namely when after each pseudo-function there are infinitely
many restoring organs in sequence. In this case, the response
to noise of the function in the executive organ can be ne-
glected, and the (infinitely long) restoring organ alone will
determine the resilience of the network. Conveniently, this
can be done with successive iterations Eq. 5, which should
eventually reach a fixed point, corresponding to the roots of
the equation (bnor/nand ◦ ǫ)2(η) = η. This is a fourth order
polynomial in η, and the roots can be obtained analytically.
The system exhibits a typical pitchfork bifurcation, with three
distinct fixed points (only two of each are stable), up to a crit-
ical value of noise pc = (3 − √7)/4 ≈ 0.0886, above which
only one fixed point exists. This corresponds to a dynami-
cal phase transition, where below this critical point the values
of the outputs will oscillate between values close to 0 and 1,
and thus memory of the past states will always be preserved.
Above the critical point, the dynamics is ergodic, independent
of any starting state. In order to characterize this transition
more precisely, we can write the expression for the previously
defined order parameter in Eq. 2 as
h = lim
t→∞
[b(t1) − b(t0)]
[8(p − pc)(p − p∗
c )]
=
1
2
(2p − 1)2
(6)
(7)
[p ≤ pc]
where b(tσ) is the value of b(t) with the starting point b(0) =
c = (3 + √7)/4. The values of the order parameter
σ, and p∗
are plotted in Fig 3. From Eq. 6 we also see easily that the
critical exponent is 1/2 (mean-field universality class). The
)
)
∞
(
b
(
b
d
η
d
1.5
1.0
0.5
0.0
0.0 0.1 0.2 0.3 0.4 0.5
0.0 0.1 0.2 0.3 0.4 0.5
p
p
)
∞
(
b
1.0
0.5
0.0
1.0
h
0.5
0.0
0.0
0.1
0.2
0.3
0.4
0.5
p
FIG. 3: Fixed points, or period-2 points, b(∞) (top) and long-term
hamming distance h (bottom) for the NAND (black) and NOR (gray)
maps, as a function of noise. Black curves correspond to NAND
restoration and grey ones to NOR.
existence of this critical value of noise points to a direct up-
3
per bound on the reliability attainable by k = 2 networks,
since it represents the maximum limit of error correction of
the restoring organ. Additionally, this critical value of noise
corresponds exactly to the upped bound found rigorously by
Evans and Pippinger [14] for reliable computation of Boolean
formulas composed of noisy NAND gates.
k > 2 -- If the functions have k > 2, the choice of the
restoring organ becomes more obvious, and the most natural
choice is the majority function, which returns simply the ma-
jority value of its inputs. Since it will work equally well if
the value on the bundle is either 0 or 1, the majority func-
tion is capable of performing restoration with only one tier of
functions, without accumulating noise in an intermediate step,
which provides it with superior characteristics. The restora-
tion map of the majority function is given by
bm(η) = 1 −
⌈k/2⌉−1
Xi=0
(cid:18)k
i(cid:19)ηi(1 − η)k−i+
δk/2,⌊k/2⌋
1
2(cid:18) k
k/2(cid:19)ηk/2(1 − η)k/2.
(8)
The last term is added only for functions with even k, which
have an indeterminate majority state.
In this case, it is as-
sumed that half the restoring functions output 1 and the other
half 0. It is also clear that majority functions with even k > 2
will perform just as well as a k − 1 odd function, and there-
fore the extra input is, for this purpose, wasted. We can anal-
yse the quality of this restoration by iterating Eqs. 8 and 3
in sequence, like it was done for k = 2. In the absence of
noise, this will lead to one of two fixed points, depending on
the starting condition. It can be seen in Fig. 4 that those fixed
points also merge into one at a critical value of noise, and the
associate order parameter h also indicates a second order tran-
sition, with the same critical exponent, but different critical
noise values. As expected, the value of h is larger for larger k,
for the same value p, and the critical noise is also larger. The
critical values pc match exactly the upper bounds for Boolean
formulas using noisy majority functions of k inputs found by
Evans and Schulman [15], given by,
pc =
1
2 −
2k−2
k−1
k(cid:0)k−1
2 (cid:1)
,
(9)
for odd k. This scales with (1/2 − pc) ∼ 1/√k (see Fig. 5),
leading to a strictly resilient situation when k → ∞.
One can test the effectiveness of the majority functions, by
considering other functions as the executive organ. Instead of
systematically analysing all 22k functions, we can consider
functions which are pathological in their stability to noise.
Here we will consider a function with maximum sensitivity,
which output 0 or 1 if all inputs are 1 or 0, respectively, but
otherwise the function is uniformly distributed, and the out-
puts will be 0 or 1 with equal probability for each input com-
bination. The corresponding map can be written as
bmin(η) =
1
2
(1 + (1 − η)k − ηk)).
(10)
Its properties can be seen in Fig. 4. We see that indeed it
becomes progressively difficult to stabilize for larger k, and
the critical point now scales as pc ∼ 1/k, and the transition
becomes first-order. On the other hand, the mere existence
of the critical point confirms some level of resilience, which
is not present in the system without redundancy, where the
critical point is always pc = 0.
)
∞
(
b
h
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
)
∞
(
b
1.0
0.8
0.6
0.4
0.2
0.0
0.00
0.25
p
0.50
0.000
0.025
p
0.05
h
1.0
0.8
0.6
0.4
0.2
0.0
k = 3
k = 4
k = 5
k = 10
k = 20
k = 100
4
shown in [16]. We stipulate that due to these robust features,
redundancy must be present in some extent in real gene reg-
ulatory networks; if not in the entire network, at least in its
more dynamically relevant modules. On the other hand, arbi-
trary redundancy close to the optimal bound is very unlikely
due to its high putative cost to the organism, which would fa-
vor instead a genetic circuit composed of fewer elements, with
only enough resilience sufficient for survival. It remains to be
seen to what extent is redundancy desirable, and how it may be
connected with other topological and functional restrictions of
gene regulation.
Acknowledgement -- I thank Tamara Mihaljev and Barbara
Drossel for carefully reviewing the manuscript, and for sug-
gestions. This work has been supported by the DFG, under
contract number Dr300/5-1.
0.00
0.25
0.50
0.000
p
0.025
p
0.05
FIG. 4: Fixed points and long-term hamming distance, for the ma-
jority restoring organ (left), and the maximum sensitivity executive
organ with the majority restoring organ (right).
c
p
−
1 2
100
10−1
10−2
10−3
10−4
10−5
c
p
10−1
10−2
10−3
10−4
10−5
10−6
10−7
10−8
100 103 106 109
100
103
106
109
k
k
FIG. 5: Critical noise value pc as a function of k, for the major-
ity restoring organ, and k = 2 NAND/NOR restoration (left), and
the maximum sensitivity executive organ with the majority restoring
organ (right). The dashed lines correspond to 1/√k (left) and 1/k
(right).
Conclusion -- We have shown that sparse networks, while
they cannot be arbitrarily resilient, they can have stable dy-
namics in the presence of noise, if redundancy is correctly
introduced. The stability is marked by second or first-order
transitions, from non-ergodic to ergodic behaviour. We obtain
upper-bounds on the error resilience attainable by redundant
networks with a given k. This is in stark contrast to what is
observed in Random Boolean Networks [10], which never ex-
hibit memory of its past states when noise is introduced, either
in its frozen or "chaotic" phases.
We have shown that the stabilization through redundancy is
successful even with the most pathologically sensitive func-
tions, such as the function with maximum sensitivity dis-
cussed. We note also that redundancy provides additional ben-
efits, such as robustness against damage and mutations, as was
∗ [email protected]
[1] R. Metzler, Physics 2, 36 (2009).
[2] S. A. Kauffman, J. Theor. Biol. 22, 437 (1969).
[3] H. H. McAdams and A. Arkin, Proc. Nat. Ac. Sci. 94, 814
(1997).
[4] E. N. Miranda and N. Parga, Europhys. Lett. 10, 293 (1989).
[5] O. Golinelli and B. Derrida, J. Phys 50, 1587 (1989).
[6] A. Aleksiejuk, J. A. Holyst, and D. Stauffer, Physica A 310,
260 (2002).
[7] Huepe and Aldana-González, J. Stat. Phys. 108, 527 (2002).
[8] X. Qu, M. Aldana, and L. P. Kadanoff, J. Stat. Phys. 109, 967
(2002).
[9] J. O. Indekeu, Physica A 333, 461 (2004).
[10] T. P. Peixoto and B. Drossel, Physical Review E (Statistical,
Nonlinear, and Soft Matter Physics) 79, 036108 (2009).
[11] B. Drossel, in Reviews of Nonlinear Dynamics and Complex-
ity, edited by H. G. Schuster (Wiley, 2008), vol. 1, ISBN
3527407294.
[12] J. V. Neumann, Automata studies p. 43 (1956).
[13] N. Pippenger, in The Legacy of John von Neumann (American
Mathematical Society, 1990).
[14] W. Evans and N. Pippenger, Information Theory, IEEE Trans-
actions on 44, 1299 (1998), ISSN 0018-9448.
[15] W. Evans and L. Schulman, IEEE Transactions on Information
Theory 49, 3094 (2003), ISSN 0018-9448.
[16] C. Gershenson, S. A. Kauffman, and I. Shmulevich, in Artifi-
cial Life X, Proceedings of the Tenth International Conference
on the Simulation and Synthesis of Living Systems, edited by
L. S. Yaeger, M. A. Bedau, D. Floreano, R. L. Goldstone, and
A. Vespignani (MIT Press, 2006), pp. 35 -- 42.
[17] We note that those distinct phases are not in general related to
the frozen or "chaotic" phases of random Boolean networks
(RBN). In the presence of noise,
the concept of frozen or
chaotic dynamics is not applicable. Furthermore, without noise,
a RBN either in the frozen or the chaotic phase can indepen-
dently be ergodic or not, according to the definition of ergodic-
ity used in this work.
[18] It is equivalent to use the AND or OR functions. We restrict
the analysis to the NAND and NOR functions without loss of
generality.
|
1608.07459 | 1 | 1608 | 2016-08-26T13:50:07 | Multiphasic interactions between nucleotides and target proteins | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.BM"
] | The nucleotides guanosine tetraphosphate (ppGpp) and guanosine pentaphosphate (pppGpp) bind to target proteins to promote bacterial survival (Corrigan et al. 2016). Thus, the binding of the nucleotides to RsgA, a GTPase, inhibits the hydrolysis of GTP. The dose response, taken to be curvilinear with respect to the logarithm of the inhibitor concentration, is instead much better (P<0.001 when the 6 experiments are combined) represented as multiphasic, with high to exceedingly high absolute r values for the straight lines, and with transitions in the form of non-contiguities (jumps). Profiles for the binding of radiolabeled nucleotides to HprT and Gmk, GTP synthesis enzymes, were, similarly, taken to be curvilinear with respect to the logarithm of the protein concentration. However, the profiles are again much better represented as multiphasic than as curvilinear (the P values range from 0.047 to <0.001 for each of the 8 experiments for binding of ppGpp and pppGpp to HprT). The binding of GTP to HprT and the binding of the three nucleotides to Gmk are also poorly represented by curvilinear profiles, but well represented by multiphasic profiles (straight and, in part, parallel lines). | physics.bio-ph | physics | Multiphasic interactions between nucleotides and target proteins
Per Nissen
Norwegian University of Life Sciences
Department of Ecology and Natural Resource Management
P. O. Box 5003, NO-1432 Ås, Norway
[email protected]
2016
2
Abstract
The nucleotides guanosine tetraphosphate (ppGpp) and guanosine pentaphosphate (pppGpp) bind to target
proteins to promote bacterial survival (Corrigan et al. 2016). Thus, the binding of the nucleotides to RsgA, a
GTPase, inhibits the hydrolysis of GTP. The dose response, taken to be curvilinear with respect to the logarithm
of the inhibitor concentration, is instead much better (P<0.001 when the 6 experiments are combined)
represented as multiphasic, with high to exceedingly high absolute r values for the straight lines, and with
transitions in the form of non-contiguities (jumps). Profiles for the binding of radiolabeled nucleotides to HprT
and Gmk, GTP synthesis enzymes, were, similarly, taken to be curvilinear with respect to the logarithm of the
protein concentration. However, the profiles are again much better represented as multiphasic than as curvilinear
(the P values range from 0.047 to <0.001 for each of the 8 experiments for binding of ppGpp and pppGpp to
HprT). The binding of GTP to HprT and the binding of the three nucleotides to Gmk are also poorly represented
by curvilinear profiles, but well represented by multiphasic profiles (straight and, in part, parallel lines).
Introduction
In addition to multiphasic profiles for ion uptake in plants (Nissen 1971, 1974, 1991, 1996), such profiles have
been recently (Nissen 2015a,b, 2016a,b,c) reported for many other processes and phenomena. In the present
paper, data for the interaction between nucleotides and target proteins in Gram-positive bacteria will be
reanalyzed to statistically compare the fits to curvilinear profiles with the fits to multiphasic profiles.
3
Reanalysis
Panels A-E and their legends not shown.
Original data kindly provided by Rebecca M. Corrigan.
Fig. 1. Authors' Fig. 2F. Quantification of the GTPase activity of RsgA in the presence of (p)ppGpp. See
also original legend.
Fig. 2 (above left). Pentaphasic profile. Transitions at -2.00, between -1.51 and -1.20 (jump), between
-0.60 and -0.30 (jump), and between 0.00 and 0.30. The absolute r value for line I is quite low (for lines
with shallow slopes, tiny errors can have large effects on the r values). High absolute r value for line III.
Lines III and IV are about parallel.
Fig. 3 (above right). Plot of deviates for the data in Fig. 2.
In addition to the r values, slopes ± SE (or only slopes) have been indicated. The probability that the better
fit to the multiphasic profile is due to chance is also given (from Fig. 3, by the Mann-Whitney rank sum
test).
4
Fig. 4 (above left). Hexaphasic profile. Transitions between -2.71 and -2.41, at -1.66, between -1.20 and
-0.90 (jump), between -0.60 and -0.30 (jump), and between 0.00 and 0.30 (single lines in the range of phases
I and II, and phases V and VI will be imprecise, with r values of -0.985 and -0.975, respectively). Very high
absolute r value for line II. Lines III and IV are about parallel.
Fig. 5 (above right). Plot of deviates for the data in Fig. 4.
. 6 (above left).
Fig. 6 (above left). Tetraphasic profile. Transitions between -2.41 and -2.11 (jump), and at -1.20 and
-0.49. Quite high absolute r value for line II, high to very high values for lines III and IV.
Fig. 7 (above right). Plot of deviates for the data in Fig. 6.
The three profiles for pppGpp differ in the number of phases (4-6). There are parallel (and adjacent)
lines in repeats 1 and 2, but not in repeat 3. There are two 3-point lines in repeat 1, one in repeat 2,
and two 3-point lines and one 4-point line in repeat 3.
5
Fig. 8 (above left). Pentaphasic profile. Transitions at -2.20, -1.51 and -0.78, and between -0.30 and 0.00
(jump). High absolute r values for lines II and III. Lines IV and V are parallel and have positive slopes.
Fig. 9 (above right). Plot of deviates for the data in Fig. 8.
Fig. 10. Hexaphasic profile. Transitions at -2.33, between -1.81 and -1.51, between -1.51 and -1.20, between
-0.90 and -0.60 (jump), and between -0.30 and 0.00 (jump). The data are insufficiently detailed in the range
of phase III for resolution of the line. Lines IV and V are parallel. There are no lines with three or more
points in the multiphasic profile, so the fits cannot be compared.
6
Fig. 11 (above left). Pentaphasic profile. Transitions at -2.18, between -1.51 and -1.20 (jump), at -0.69, and
between -0.30 and 0.00 (jump). Exceedingly high absolute r value for line II. Lines IV and V are parallel.
Fig. 12 (above right). Plot of deviates for the data in Fig. 11.
As also for pppGpp, the three profiles for ppGpp differ in the number of phases (5 or 6) and in the
number of 3-point lines (0-2). However, there is a set of adjacent and parallel lines in each of the profiles.
In summary, the data in Fig. 2F are well represented by multiphasic profiles. The transitions between
adjacent and parallel lines are necessarily in the form of noncontiguities (jumps), and the data should not be
represented by curvilinear profiles. The finding of high to exceedingly high absolute r values for the straight
lines also shows that the profiles are multiphasic rather than curvilinear. P<0.001 that the better fit to
multiphasic profiles is due to chance (by Fisher's (1954) method for combining independent probabilities).
7
Fig. 13. Authors' Fig. S2A. Binding curves for radiolabeled ppGpp, pppGpp, and GTP with purified
HprTSA. For clarity, the points above are also shown in the plots below.
Plots for HprT, ppGpp
Fig. 14 (above left). Heptaphasic profile. Transitions at -1.33 and -0.48, between -0.11 and 0.19, between
0.19 and 0.49, between 0.80 and 1.10 (jump), and between 1.40 and 1.70. The data are insufficiently
detailed in the range of phase IV for resolution of the line. Line I is horizontal, line II has a high r value, and
lines V and VI are about parallel. The pattern for phases IV-VII is identical to the pattern in the same range
for repeat 1 for pppGpp (Fig. 22).
Fig. 15 (above right). Plot of deviates for the data in Fig. 14.
8
Fig. 16 (above left). Hexaphasic profile. Transitions between -1.91 and -1.61 (jump), between -0.71 and
-0.41 (jump), between -0.11 and 0.19 (jump), between 0.49 and 0.80 (jump), and between 1.10 and 1.40
(jump). Low r value for line I, but shallow slope. Quite high r value for line II.
Fig. 17 (above right). Plot of deviates for the data in Fig. 16.
Fig. 18 (above left). Hexaphasic profile. Transitions between -2.21 and -1.91 (jump), at -1.43 and -0.01,
between 0.49 and 0.80, and between 0.80 and 1.10. The data are insufficiently detailed in the range of phase
V for resolution of the line. High r values for the 5-point line III and for line VI. Lines I and II are parallel
and have slightly negative slopes.
Fig. 19 (above right). Plot of deviates for the data in Fig. 18.
9
Fig. 20 (above left). Hexaphasic profile. Transitions between -1.91 and -1.61 (jump), between -1.01 and
-0.71 (jump), between -0.11 and 0.19 (jump), between 0.49 and 0.80 (jump), and between 1.10 and 1.40
(jump). Low r value for line I, but shallow slope. High r values for lines II and III. Lines III and IV are
parallel, as are lines V and VI.
Fig. 21 (above right). Plot of deviates for the data in Fig. 20.
Characteristics of multiphasic profiles for HprT, ppGpp:
Heptaphasic profile for repeat 1, hexaphasic profiles for repeats 2-4. One set of adjacent and parallel lines
for repeats 1 and 3, two sets for repeat 4. Lines I and II in repeat 3 have negative slopes.
Plots for HprT, pppGpp
Fig. 22 (above left). Octaphasic profile. Transitions between -2.21 and -1.91 (jump), between -1.61 and
-1.31 (jump), at -0.71, between -0.11 and 0.19, between 0.19 and 0.49, between 0.80 and 1.10 (jump), and
between 1.40 and 1.70. The data are insufficiently detailed in the range of phase V for the line to be
resolved. A single line in the range of phases I-III will have a low r value (0.790). High r values for lines III
and IV. Line I has a slightly negative slope.
Fig. 23 (above right). Plot of deviates for the data in Fig. 22.
10
Fig. 24 (above left). Heptaphasic profile. Transitions between -2.52 and -2.21, between -1.61 and -1.31
(jump), between -1.01 and -0.71, between -0.71 and -0.41, between 0.49 and 0.80 (jump), and at 1.12. The
data are insufficiently detailed in the range of phase IV for the line to be resolved. A single line in the range
of phases I-IV will have a low r value (0.869). The 4-point line V has a high r value. Line VII is horizontal.
Lines II and III are parallel and have slightly negative slopes. Lines V and VI are also about parallel.
Fig. 25 (above right). Plot of deviates for the data in Fig. 24.
Fig. 26 (above left). Octaphasic profile. Transitions between -2.21 and -1.91 (jump), between -1.61 and
-1.31, between -1.31 and -1.01, between -0.41 and -0.11 (jump), between 0.19 and 0.49, between 0.49 and
0.80, and between 1.40 and 1.70. The data are insufficiently detailed in the range of phases III and VI for the
lines to be resolved. A single line in the range of phases I-III will have r = -0.430. Exceedingly high r value
for line IV. Line VII has a markedly negative slope and a quite high absolute r value.
Fig. 27 (above right). Plot of deviates for the data in Fig. 26.
11
Fig. 28 (above left). Hexaphasic profile. Transitions at -1.91, between -1.31 and -1.01 (jump), at 0.19,
between 0.80 and 1.10 (jump), and between 1.40 and 1.70. Line I has a markedly negative slope and a quite
high absolute r value. The r values of line II and the 5-point line III are quite high and high, respectively.
The r value of line IV is quite low, but the 3-point line is parallel with line V.
Fig. 29 (above right). Plot of deviates for the data in Fig. 28.
Characteristics of multiphasic profiles for HprT, pppGpp:
Octaphasic profiles for repeats 1 and 3, heptaphasic profile for repeat 2, hexaphasic profile for repeat 4. Two
sets of adjacent and parallel lines for repeat 2, one set for repeat 4. Line I in repeat 1, lines I and VII in
repeat 2, line VII in repeat 3, and line I in repeat 4 have negative slopes.
Plots for HprT, GTP
The fits to the curvilinear profiles are clearly much poorer than the fits to the multiphasic profiles, and P
values for the significance of this difference have not been calculated.
Fig. 30 (above left). Pentaphasic profile. Transitions between -2.21 and -1.91 (jump), between -1.61 and
-1.31 (jump), between -0.71 and -0.41 (jump), and between 0.19 and 0.49 (jump). High r values for lines III,
IV and the 5-point line V. Lines III and IV are parallel.
Fig. 31 (above right). Octaphasic profile. Transitions at -1.95, between -1.61 and -1.31 (jump), between
-1.01 and -0.71 (jump), between -0.41 and -0.11 (jump), between 0.19 and 0.49 (jump), between 0.80 and
1.10 (jump), and between 1.40 and 1.70. Lines III and IV are parallel, as are lines V and VI.
12
Fig. 32 (above left). Hexaphasic profile. Transitions between -1.91 and -1.61 (jump), at -0.87 and 0.42,
between 0.80 and 1.10 (jump), and between 1.40 and 1.70. High and very high r values for lines I and III,
quite low absolute r value for line II which has a markedly negative slope. Lines IV and V are parallel. The
value for the point at -0.41 is probably in error (see the very good fit to line III) and has been omitted from
the calculations.
Fig. 33 (above right). Heptaphasic profile. Transitions between -2.21 and -1.91 (jump), between -1.31 and
-1.01 (jump), between -0.71 and -0.41 (jump), between -0.11 and 0.19 (jump), between 0.49 and 0.80, and
between 0.80 and 1.10. The data are insufficiently detailed in the range of phase VI for resolution of the line.
High and very high r values for lines II and VII. Lines III and IV are parallel.
Characteristics of multiphasic profiles for HprT, GTP:
The profiles are octaphasic, heptaphasic, hexaphasic and pentaphasic for repeat 2, 4, 3 and 1, respectively.
Two sets of adjacent and parallel lines for repeat 2, one set for each of the other repeats. Negative slope for
line II in repeat 3.
13
Conclusion
From a comparison of fits it is clear, at high levels of significance, that the present data cannot be acceptably
represented by curvilinear profiles. They are, instead, very well represented by multiphasic profiles, i.e. by
profiles consisting of a series of straight lines separated by discontinuous transitions.
Acknowledgment – I am very grateful to Bob Eisenberg for his continued interest and encouragement.
References
Corrigan RM, Bellows LE, Wood A, Gründling A (2016) ppGpp negatively impacts ribosome assembly
affecting growth and antimicrobial tolerance in Gram-positive bacteria. Proc Natl Acad Sci USA E1710-E1719.
Fisher RA (1954) Statistical Methods for Research Workers, 12th ed., section 21.1, Oliver Boyd, Edinburgh.
Nissen P (1971) Uptake of sulfate by roots and leaf slices of barley. Physiol Plant 24: 315-324.
Nissen P (1974) Uptake mechanisms: Inorganic and organic. Annu Rev Plant Physiol 25: 53-79.
Nissen P (1991) Multiphasic uptake mechanisms in plants. Int Rev Cytol 126: 89-134.
Nissen P (1996) Uptake mechanisms. Pp. 511-527 in: Waisel Y, Eshel A, Kafkafi U (eds.). Plant Roots. The
Hidden Half (2. ed.). Marcel Dekker, Inc. New York.
Nissen P (2015a) Discontinuous transitions: Multiphasic profiles for channels, binding, pH, folding and chain
length. Posted on arXiv.org with Paper ID arXiv:1511.06601.
Nissen P (2015b) Multiphasic pH profiles for the reaction of tris-(hydroxymethyl)-aminomethane with phenyl
esters. Posted on arXiv.org with Paper ID arXiv:1512.02561.
Nissen P (2016a) Profiles for voltage-activated currents are multiphasic, not curvilinear. Posted on arXiv.org
with Paper ID arXiv:1603.05144.
Nissen P (2016b) Multiphasic profiles for voltage-dependent K+ channels: Reanalysis of data of MacKinnon
and coworkers. Posted on arXiv.org with Paper ID arXiv:1606.02977.
Nissen P (2016c) 'Perfectly' curvilinear profiles for binding as determined by ITC may in fact be multiphasic.
Posted on arXiv.org with Paper ID arXiv:1606.09133.
|
1310.1483 | 1 | 1310 | 2013-10-05T15:06:28 | Aggregation of Red Blood Cells: From Rouleaux to Clot Formation | [
"physics.bio-ph",
"cond-mat.soft",
"q-bio.CB"
] | Red blood cells are known to form aggregates in the form of rouleaux. This aggregation process is believed to be reversible, but there is still no full understanding on the binding mechanism. There are at least two competing models, based either on bridging or on depletion. We review recent experimental results on the single cell level and theoretical analyses of the depletion model and of the influence of the cell shape on the binding strength. Another important aggregation mechanism is caused by activation of platelets. This leads to clot formation which is life saving in the case of wound healing but also a major cause of death in the case of a thrombus induced stroke. We review historical and recent results on the participation of red blood cells in clot formation. | physics.bio-ph | physics | Aggregation of Red Blood Cells: From Rouleaux to Clot Formation
C. Wagner a , P. Steffen a , S. Svetina b
a Experimentalphysik, Universit at des Saarlandes, Postfach 151150, 66041 Saarbr ucken, Germany
b Institute of Biophysics, Faculty of Medicine, University of Ljubljana, and Jo zef Stefan Institute, Ljubljana, Slove nia
3
1
0
2
t
c
O
5
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
3
8
4
1
.
0
1
3
1
:
v
i
X
r
a
Abstract
Red blood cells are known to form aggregates in the form of rou leaux. This aggregation process is believed to be
reversible, but there is still no full understanding on the b inding mechanism. There are at least two competing models,
based either on bridging or on depletion. We review recent ex perimental results on the single cell level and theoretical
analyses of the depletion model and of the in fluence of the cel l shape on the binding strength. Another important
aggregation mechanism is caused by activation of platelets. This leads to clot formation which is life saving in the
case of wound healing but also a major cause of death in the case of a thrombus induced stroke. We review historical
and recent results on the participation of red blood cells in clot formation.
R ´esum ´e
Il est bien connu que les globules rouges forment des agr ´ega ts, connus sous le nom de rouleaux . Il est souvent
admis que ce ph ´enom `ene d’agr ´egation est r ´eversible, ma is l’ ´elucidation pr ecise des m ´ecanismes `a l’ oeuvre dans
le
processus conduisant `a la liaison entre globules rouges es t loin d’ etre achev ´e. Il existe dans la litt ´erature au moi ns deux
mod `eles distincts, l’un est bas ´e sur la formation de ponts mol ´eculaires et l’autre sur la notion de d ´epltion. Nous pa ssons
en revue les r ´esultats exp ´erimentaux r ´ecents l’ ´echell e cellulaire et analysons le modle th ´eorique bas ´e sur la no tion de
d ´epltion. Nous discuterons l’in fluence de la forme cellula ire sur la force de liaison. Un autre m ´ecanisme d’agr ´egati on
jouant un r ole important in vivo est celui associ ´e l’activ ation plaquettaire. Ceci peut conduire la formation de cail lots
sanguins, processus vital lorsqu’il s’agit de cicatrisation de blessures, mais qui peut etre ´egalement fatal, const ituant
une cause majeure de d ´ecs, lorsqu’il s’agit de thrombose.
Keywords:
red blood cells, depletion, aggregation
mots-cl ´es: globules rouges, d ´epletion, aggr ´egation
1. Introduction
There are at least three major classes of aggregation of red b lood cells (RBCs). First the so-called rouleaux
formation that is caused by the plasma macromolecules and that is supposed to be a reversible aggregation process,
second there are several indications that RBCs might become actively adhesive in the presence of a blood clot. And
finally there are many pathological cases like malaria or sic kle cell diseases where red blood cells are known to form
large aggregates that hinder the flow of blood [1, 2, 3]. This c ontribution will cover the first two of these three cases.
In a recent book [4] Baskurt et al. gave a comprehensive overv iew on the first topic. Here we intend to express
some fresh views about the interplay between the effects on the aggregation and coagulation of the solution and of the
intrinsic properties of RBCs. Most of the recent findings are largely based on the latest progresses in the application
and analyses of single cell measurements.
Human RBCs in blood samples of healthy donors have a tendency to form aggregates that look similar to a stack
of coins [4, 5, 6, 7]. These linear structures are called roul eaux. Figure 1a shows a single rouleau in autologous plasma
consisting of seven individual RBCs. The number of RBCs per rouleau can vary and branching into two rouleaux can
occur. The picture of the aggregate shown in Fig. 1a was taken under static conditions meaning that no flow was
applied to the sample. The attractive forces involved in the creation process of rouleaux are relatively weak. Hence,
it is possible to dissolve rouleaux into smaller fractions o r even into single cells by applying a sufficient shear force
[8, 9]. This leads to the pronounced shear thinning of blood. At low shear rates large aggregates lead to an increased
Preprint submitted to Compte rendues
October 8, 2013
viscosity, but by increasing the shear rate the aggregates b reak up and the viscosity decreases to a constant high shear
rate value. For RBCs without macromolecules and thus no rouleaux formation shear thinning is much less pronounced
and results from deformation and orientation of the RBCs only [10]. There exist two competing theories to explain
the aggregation mechanism. They are based either on a model o f bridging or on depletion (Fig. 1b and c).
Figure 1: a) From ref. [37]. Snapshot of a rouleaux of 7 RBCs in a dextran solution. b) Schematic illustration of how macromolecular bridging
leads to intercellular adhesion. Macromolecules adsorb on to the RBC-membrane and are able to bridge the adjacent cell. c) Illustration of the
depletion phenomenon in a binary colloidal system. The depletion layer of the colloids with diameter d1 (orange) are marked white. When two
depletion layers overlap, a volume δV > 0 (dark blue) is released, which is additionally available for the smaller macromolecules (green) with
diameter d2 .
Formation of rouleaux is also affected by the intrinsic properties of RBCs such as the elastic behavior of the RBC
membrane which contributes to the resistance of RBCs to aggregate. It is therefore of crucial importance to understand
the interplay between membrane elastic energy and shapes of the agglomerate.
In general, RBCs can pass capillaries with diameters smaller than their cell size, but if clotting occurs, the flow
might come to a complete stop. In the case of wound healing, clotting is life-saving, but in a healthy vessel a thrombus
might lead to a stroke, the main cause of death in the developed world. The blood coagulation process is a complex
process that involves many collective players and factors. In a brief manner, the coagulation process starts with the
activation of blood platelets (e.g. at a damaged vessel wall). The activated platelets release a variety of messengers
and growth factors in order to recruit other cells to partici pate in blood clot formation. Obviously, RBCs are a major
part of the thrombus, but it is commonly believed that they are simply trapped in the fibrin network due to their
prevalence in the blood [11]. Therefore, the role of RBCs is a lways assumed to be completely passive. As early
as a century ago, the first clinical studies described a corre lation between a decreased concentration of RBCs, i.e.
haematocrit value, and longer bleeding times [12]. These early results have been con firmed by a number of clinical
investigations [13, 14]. Considering the high percentage of RBCs in the blood and the clinical indications, it is evident
that a profound understanding of the role of RBCs in clotting is crucial. An adhesion of RBCs to platelets – and not to
each other – was assumed to be of principal importance [14]. A co-adhesion of healthy RBCs was not assumed until
recently Kaestner et al. [15] suggested a signaling cascade that predicts an active participation of RBCs in blood clot
formation. In first studies [16, 17] the active co-adhesion o f RBCs could be shown, but statistically significant and
quantitative data are rare.
2
2. Rouleaux formation
Rouleaux are caused by the presence of macromolecules, such as fibrinogen in blood plasma. In both, coagulation
and aggregation the fibrinogen plays a crucial role [4]. In co agulation it is converted into fibrin that is polymerized
to form a mesh in which platelets and RBCs are trapped [18]. Similar rouleaux formation can also be induced by re-
suspending the RBCs in physiological solutions containing neutral macromolecules such as dextran [19]. However,
without any macromolecules, e.g. RBCs in a simple salt solution, no aggregation occurs. The fibrinogen mediated ag-
gregation of RBCs increases consistently with increasing fi brinogen concentration [20], whereas the dextran-mediated
aggregation of RBCs reaches a maximum at a certain dextran concentration. The strength of the aggregation depends
not only on the dextran concentration, but also on the molecu lar weight of the dextran (i.e., the radius of gyration of
the dextran) [6, 21]. To this day the mechanisms involved in RBC aggregation have not been fully understood. There
are two coexisting models that try to explain the rouleaux fo rmation of RBCs: the bridging model and the depletion
model. In the bridging model, it is assumed that fibrinogen or dextran molecules non-specifically adsorb onto the cell
membrane and form a ”bridge ” to the adjacent cell [21]. In con trast, the depletion model proposes the opposite. In
this model, aggregation occurs because the concentration o f macromolecules near a RBC surfaces in close proximity
is depleted compared to the concentration of the bulk phase, resulting in a net ”depletion ” force.
The bridging model assumes that large macromolecules adsorb onto the cell surface and thereby bridge two ad-
jacent cells. When these bridging forces exceed the disaggregation forces such as electrostatic repulsion, membrane
strain and mechanical shearing, aggregation occurs [6, 21, 22, 23, 24]. More precisely, studies that focused on the
inter cellular distance [5] of two adjacent cells showed that the intercellular distance is less than the size of the hy-
drated molecules. This leads to the assumption that the terminal portions of the flexible polymers are adsorbed onto
the surfaces of adjacent cells, resulting in a cell-cell adhesion (see Fig. 1). Thereby, the cell-cell distance increases
with increasing polymer size but is always smaller than the d iameter of the hydrated polymer [25].
In the depletion model aggregation occurs because the concentration of macromolecules near a RBC surface in
close proximity is depleted compared to the concentration o f the bulk phase, resulting in a net ”‘depletion ”’ force.
A first explanation of depletion forces was given by Asakura a nd Oosawa [26], who discovered that the presence of
small spheres (i.e., macromolecules) can induce effective forces between two larger particles if the distance b etween
them is small enough. The origin of these forces is purely entropic. When two large plates are immersed in a solution
of rigid spherical macromolecules and the distance between the inner surfaces of these two plates is smaller than the
diameter of solute macromolecules, none of these macromolecules can enter the space between the plates and this
space becomes a phase of the pure solvent. Therefore, a force equivalent to the osmotic pressure of the solution of
macromolecules acts on the outer surfaces of these plates. Such a force also appears between two spherical particles
if the distance between the two large particles decreases to less than the size of the surrounding macromolecules
(Fig. 1c). In such a system a so-called depletion layer surrounds the large particles where the centers of the smaller
particles cannot enter. Consistently, in that depletion layer the concentration of macromolecules becomes depleted
compared to that of the bulk. The thickness of this depletion layer equals the radius of the smaller particles. When
overlapping between two depletion layers occurs, an additional free volume is available for the smaller particles
causing an increase in entropy and hence a decrease in Helmho ltz’s free energy leading to an effective osmotic pressure
causing an attractive force between the large particles.
In diluted solutions below the overlap concentration [27]
polymers can be treated as rigid spheres with a radius d2 /2 matching the radius of gyration of the polymer [28].
2.1. Bridging versus depletion
Over the past, there have been studies in support of both theo ries. Studies in favor of the bridging model dealt with
either aggregation induced by nonspecific binding of macrom olecules [29, 25] or by specific binding mechanisms [30].
The determination of macromolecular adsorption of polymer s and proteins to RBCs are subject to a lot of possible
artifacts and consequently the interpretation of existing data is difficult [31, 32]. Despite a lot of efforts to quantify
macromolecular binding, conclusive data is still lacking. On the other side, several studies favoring the depletion
model have been published [33, 34, 35]. Neu and Meisselman [35] adopted the depletion concept and applied it on
the aggregation of red blood cells and developed a theoretical description of the acting forces. They also predicted
quantitative values of the interaction energies.
3
2.2. Theoretical description of depletion based aggregation of RBCs
The nature of the interaction forces depends on the surfaces of the adhering objects. The surface of the RBCs
is strongly in fluenced by the soft layer attached to the membr ane, called glycocalyx. Therefore, one has to take the
possibility into account that the RBCs also interact via steric interaction due to the overlapping glycocalyces. Due to
the high electrostatic repulsion, cell-cell distances at which minimal interaction energies occur (i.e. maximal adhesion
strength) are always greater than twice the thickness of the glycocalyx [35]. Thus, steric interactions can be neglected
for the case of RBCs and only depletion and electrostatic repulsion have to be considered. Neu and Meisselman [35]
computed the effects of bulk phase polymer concentration on interaction ene rgy. They identified the softness and the
consecutive penetration of dextran molecules into the soft RBC surface as crucial when it comes to the development of
the characteristic bell-shaped relations of the interaction energy in dependence on the dextran concentration (Fig. 2)
while without surface penetration one finds a linear depende nce of the interaction energy on the dextran concentration.
At first sight these bell-shaped curves might seem counterin tuitive because considering depletion interaction one
expects a consecutive increase in the interaction energy wi th increasing polymer concentration. The reason for these
bell-shape nature of the curves lies in the increasing penetration depth of the polymers into the glycocalyx with
increasing bulk polymer concentration and in the decreasing depletion layer thickness with increasing bulk polymer
concentration. In order to compare the computed model with experimental data, Neu and Meisselman [35] took data
from Buxbaum et al. [36]. There, a micro pipette based approach was chosen to measure the surface affinities of
a RBC membrane vesicle with an intact RBC. Neu and Meisselman [35] varied the penetration constant until the
calculated peak interaction energy for dextran 70 kDa or dex tran 150 kDa equaled the value reported by Buxbaum et
al. [36].
Figure 2: From ref. [37]. a) The measured adhesion forces for different concentrations of dextran (circles: dextran 70, tria ngles: dextran 150). The
maximum interaction strengths were observed at 2g/dl (dextran 70) and 4g/dl (dextran 150). b) Dependence of the interaction energy of two red
blood cells for different concentrations of dextran. The solid line represents the curve calculated by Neu and Meisselman [35].
2.3. Single cell force measurements
The technique of single cell force spectroscopy (SCFS) was used to measure the interaction energies between
human red blood cells as functions of the molecular weight and concentration of dextran [37]. The dextrans used
were dextran70 (DEX70 with a molecular weight of 70 kDa), and dextran150 (DEX150 with a molecular weight of
150 kDa) from Sigma-Aldrich. The measurements were conducted at the single cell level and were compared to the
predicted values of Neu and Meisselman [35]. An atomic force microscope (AFM) (Nanowizard 2, equipped with the
CellHesion Module with an increased pulling range of up to 100 µm, JPK Instruments, Germany) was used to conduct
single cell force spectroscopy measurements [38]. In the course of the experiment, a single RBC was attached to an
4
AFM cantilever by appropriate functionalization. Cell TakT M (BD Science) was used to bind a cell to the cantilever.
After cell capture, the cantilever was lowered onto another cell, and the adhesion force and adhesion energy were
measured. The retraction curve is typically characterized by the maximum force required to separate the cells from
each other and adhesion energies are calculated by computing the area under the retraction curve of the force distance
curve. The interaction energies (more precisely, the inter action energy densities of two RBCs) are calculated by
dividing the measured adhesion energies by the contact areas of the adhering cells using a value of 50.24 µm2 derived
from the maximum radius of RBCs. Figure 2a and b show the dependence of the adhesion force and the interaction
energy on the concentration of the dextran used.
Since little statements regarding adhesion forces are made by the literature, it is important to consider the interac-
tion energies, too. The measured values of the adhesion energy can now be compared to the values predicted by the
theory [35]. The excellent agreement of the SCFS measuremen ts with the theoretical description give more rise to the
assumption that the driving force in rouleaux formation is r ather depletion induced than bridging induced. As could
be seen in contact time dependence measurements, the adhesion energies increase with increasing contact time. This
could be due to bridging, but there is not enough data to conclusively decide that. With the present data it appears that
the rouleaux formation, at least at the beginning, is purely depletion mediated.
2.4. The role of RBC shape transformations in the RBC aggregation processes
Aggregated RBCs exert spatial constraints on each other and therefore their shapes differ from the shapes that they
attain when free. As the shape of the unconstrained, free cel l corresponds to the smallest possible membrane elastic
energy, in an aggregated RBC this energy is larger. RBC aggregation can thus only occur if the total energy decrease
due to the attraction between cells exceeds the corresponding increase of the elastic energy of their membranes. The
knowledge about the factors that affect RBC membrane elastic energy is therefore for the underst anding of the RBC
aggregation process equally important as the knowledge about the factors involved in cell-cell interactions. In this
subsection we shall discuss the relationship between RBC shapes in RBC aggregation process at the macroscopic
level. In order to reveal the essential features of the role o f RBC shapes in the aggregation process, we shall give an
account mainly on the works on the simplest possible (minimal) model of the RBC aggregation. Within this model
the RBC is assumed to behave analogously to the behavior of lipid vesicles with homogeneous bilayer membranes,
therefore we shall review also some works on the adhesion of vesicles. When applicable we shall also comment on
the limitations of the treated minimal model.
The general feature of the adhered RBC is that its membrane is divided into zones which are in contact with
surfaces of other cells and zones which are in contact only wi th the surrounding solution. The adhesion energy
pertains only to the zones that are in contact. In the minimal-prototype model of the RBC aggregation process it is
assumed that membranes involved are laterally homogeneous and that it is possible to de fine an adhesion energy ( Wa )
which is the product of the contact area (Ac ) and an adhesion constant (Γ), e.g. measured in J/µm2 (see Fig. 2):
Wa = −ΓAc
(1)
Shapes of adhered or aggregated RBCs correspond to the minimum of the energy functional which in addition to Eq.
1 involves the elastic energy of its membrane. The latter is the sum of the elastic energy of the bilayer part of the RBC
membrane and the elastic energy of its membrane skeleton. The minimal model of the RBC aggregation takes into
consideration only the local [39] and non-local [40, 41, 42, 43] bending energy terms of the RBC membrane bilayer,
expressed, respectively, as
1
2
1
2
kr
h2A0
Wb =
(∆A − ∆A0 )2
kc I (C1 + C2 − C0 )2dA +
where kc is the bending modulus of the bilayer, C1 and C2 are the principal curvatures of the membrane, and C0 is
the spontaneous (preferred) curvature of the membrane. kr is the nonlocal bending modulus, ∆A is the difference
between the areas of the outer and the inner lea flets equal to h H (C1 + C2 )dA where h is the distance between neutral
surfaces of the membrane lea flets, ∆A0 is the equilibrium (preferred) difference between the areas of the outer and
the inner lea flets which essentially depends on the di fference in the molecular occupancy of the two layers, and A0
is the equilibrium (preferred) membrane area. The value of t he local bending constant for the RBC membrane is
kc = 2 x10−19 J while kr is about twice that large [44]. If the lateral tensions of the two bilayer lea flets equilibrate,
5
(2)
e.g. by way of the flip- flop transport of lipid molecules, the n
on-local bending term vanishes. The general problem
to be solved in the RBC aggregation phenomenon is to find the mi nimum of the sum of Eqs. 1 and 2 for all cells
that form a given aggregate. Technically the problem is a generalization of the shape determination of a single
unconstrained vesicle/cell which corresponds to the minimization of Eq. 2. The latter topic has been in the past
reviewed comprehensively [45, 46]. Shapes can be theoretically determined by solving the corresponding shape
equation which gives beside the coordinates of all membrane points also the corresponding principal curvatures. For
the present discussion it is of interest to note that at the reduced volume (volume divided by the volume of a sphere
with the same membrane area) of the RBC which is about 0.6, the shape that corresponds to the minimum of the
local bending energy is a discocyte and thus coincides with the shape of the RBC in its resting state. The decrease
of either spontaneous curvature C0 , equilibrium area difference ∆A0 , or both causes RBC to transform into an oblate
stomatocyte. The increase of these quantities tends to make the shapes prolate, however, RBCs instead transform into
spiculated echinocytes which has been interpreted by taking into account the elasticity of the RBC membrane skeleton
[47]. This means that the minimal model of the RBC aggregation is applicable only when RBC shape transformations
do not involve too strong changes of membrane principal curvatures.
When measuring the relative effects of the adhesion energy it is appropriate to look into how its magnitude com-
pares to the membrane bending energy. A suitable parameter t hat was chosen to quantify this comparison is the ratio
between the adhesion energy corresponding to the area of the cell membrane Γ4πR2
0 (where R0 is the radius of the
sphere with the RBC membrane area A0 ) and the bending energy of a spherical membrane (8πkc),
γ = ΓR2
0 /2kc .
(3)
For γ ≪ 1 the adhesion is negligible and for γ ≫ 1 it prevails. Eq. 3 also indicates that at given values of the adhesion
constant and the reduced volume the effects of the adhesion are more pronounced in larger cells.
In treating the RBC shape behavior in the aggregation processes it is necessary to take into account that we are
dealing with the adhesion that occurs between two flexible su rfaces. This is important to note because most works
on vesicle and cell adhesion deal with their adhesion to rigid surfaces. When RBC is adhered to a rigid surface its
contact zone attains the shape of the surface. It is then only necessary to determine the shape of the non-adhering
membrane zones which can be done by solving the shape equation for those membrane sections. The variables in
the corresponding shape determination are the membrane pri ncipal curvatures and the area of the adhesion zone. For
solving the shape equation it is necessary to know the values of principal curvatures at the zone boundaries. For the
vesicle adhesion to the flat rigid surface this condition was derived [48] and reads for axisymmetrical shapes
cm = (4γ)1/2,
(4)
where cm is the reduced value of the principal curvature along the mer idians, cm = R0Cm . When the adhesion occurs
between two flexible surfaces which is the case in the RBC-RBC adhesion, the shape equation has to be solved for
all membrane zones. For two axisymmetric adhered cells with membranes of equal bending constants the generalized
boundary conditions read [49]
∆cm = (2γ)1/2 ,
where ∆cm is the difference between the reduced principal curvature along merid ians of the free and adhered parts of
the membrane. It was shown that also for vesicles of a general symmetry the squared curvature jump 4γ demanded
by the rigid surface version (Eq. 4) is shared in equal parts ( 2γ) between the two membranes [50]. The problem of
boundary conditions for adhered soft membranes was recentl y also comprehensively treated by Agrawal [51].
The classical example of RBC aggregation is the formation of the rouleau. Rouleaux were first treated by the
minimal model by Skalak et al. [52]. They considered the adhesion zone to be flat. Consequently, each cell in such
a linear array was symmetrical with respect to the mirror equ atorial plane. More recently the question was raised
if the cells in the rouleau can have no such symmetry [49]. Rouleau was treated as an in finite chain of identical
axisymmetrical cells so that it was possible to apply period ic boundary conditions and determine the RBC shape by
treating a single cell. It was shown that there is the transit ion from mirror symmetrical to asymmetrical shapes that
occurred by lowering the reduced volume or /and by increasing the adhesion constant. The shape parameter that favors
the asymmetrical shapes was shown to be also the preferred difference between the areas of the bilayer layers (∆A0 ).
The question about possible symmetry of adhered RBCs was reexamined by applying numerical methods for RBC
(5)
6
shape determination and thus removing the restriction abou t the axial symmetry [53]. RBC doublet was studied and
it was shown that the RBCs in a doublet indeed exhibit a transition from axially symmetrical to non-axisymmetrical
shapes. When the reduced adhesion strength γ is small but larger than the threshold for adhesion, the contact zone
is planar and circular as if each of the vesicles would stick to the rigid surface, and the shapes of the two vesicles
are the same. If the adhesion strength is increased further, the stable doublet consists of identical vesicles joined in
a sigmoidal, S-shaped contact zone with an invagination and a complementary evagination on each vesicle. There
are two features of these doublets which may have a role in the formation and the outlook of rouleaux. At some
intermediate adhesion strengths the membranes of the two constituent cells of a doublet are in the contact zones
rather curved, whereas their outer surfaces are practicall y flat. Such doublets could adhere to each other without an
additional increase of the membrane elastic energy. Because the cells in a doublet with a sigmoidal contact zone are
shifted away from the rouleau long axis in opposite directions, in the so formed rouleaux the cells arrange in a zig-zag
manner. Such cell arrangements have been indeed observed [52, 54].
The study of doublet shapes on the basis of the described minimal model of RBC aggregation has been employed
to envisage different adhesion regimes with respect to the reduced adhesion strength γ and characterized by a specific
type of aggregate [55]. In the regime right above the adhesion threshold (0.3 . γ . 2) the predominant aggregate
shape is the flat-contact doublet. In the weak adhesion regim e ( 2 . γ . 4) doublets aggregate into the zig-zag
rouleaux while in the strong adhesion regime ( 6 . γ . 10) predominate sigmoid-contact round doublets. At still
higher values of the adhesion parameter (γ & 15) the cells would favor to aggregate into rounded clumps. In general
it looks that rouleaux are formed in the regime in which there is still a reasonable balance between the adhesion and
bending energies.
Because of many limitations of the minimal model for RBC aggregation the above conclusions are only qualitative.
First of all, the minimal model does not consider the area expansivity and shear energy terms of RBC membrane
skeleton [56, 57] which, as already stated, becomes important especially in shape transformations to or from the
shapes exhibiting high curvatures. It can be concluded that the aggregation behavior depends on the initial RBC
shapes and is different if they are for example discocytes or echinocytes. The next possible future generalization of
the minimal model is the effect of adhesion on the lateral segregation of membrane components. In the aggregation
where RBCs adhere to other blood cells, it will be important to derive the boundary conditions analogous to Eq.
5 which would take into consideration that the two adhering membranes have different mechanical properties [51].
There are also still unanswered problems related to the size and structure of rouleaux. RBCs are variable in their sizes
and this certainly affects the sizes of rouleaux, especially it is expected that the end cells could have very specific sizes.
The effects of the initial RBC shapes are of particular interest because non-discoid shapes have been found in some
hereditary or other diseases. The observation of RBC shapes and their aggregation behavior may serve as possible
indicators of these diseases.
3. Platelet induced coagulation of red blood cells
Coagulation is governed by a complex signaling cascade and it requires a number of essential enzymes and cofac-
tors, so-called coagulation factors, and is finalized by the formation of the thrombus [58]. The platelets themselves are
the most important cells in the coagulation process, inside those platelets the so-called scramblase protein is activated.
The scramblase activation results in a degradation of the asymmetrical distribution of phospholipids in the lipid bilayer
of the RBC membrane. This leads to an exposition of the negatively charged phospholipid phosphatidylserine (PS)
in the outer lea flet of the RBC membrane. This PS is under suspi cion to play an important role in blood coagulation
since it might provide a catalytic surface for prothrombina se complexes [59]. Much work has already been conducted
on the hypothesis of an active participation of RBCs in thrombus formation, but a direct investigation of the involved
adhesion forces in order to check this hypothesis was missing until recently. Optical tweezers as well as single cell
force spectroscopy were used to check this hypothesis and to quantify the occurring adhesion in terms of adhesion
strength [17].
The activation of RBCs by activated platelets involves a specific signaling cascade (Fig. 3). Kaestner and Bern-
hardt hypothesized a Ca2+ in flux via a non-selective cation channel (NSC) [60, 61, 15] w hich is permeable to different
mono and bivalent cations such as N a+ , K + or also Ca2+ . This channel is opened by physiological concentrations of
prostaglandin E2 (PGE2 ) and lysophosphatidic acid (LPA) [62, 15, 63]. These substances are released by activated
platelets. The increased intracellular Ca2+ -concentration acts as a trigger mainly for two processes. F irst, another
7
Figure 3: Signaling Cascade of the LPA induced adhesion of RBCs.
calcium-dependent channel (the Gardos channel) is opened by the increased calcium level [64]. Through this channel,
intracellular potassium effluxes out of the cell, followed by a shrinkage of the cell [65, 66]. The second, more impor-
tant, process is that the lipid scramblase protein is activa ted which has a profound consequence: the breakdown of the
asymmetrical lipid distribution between both lea flets [67, 68, 69, 70, 71]. In this way, the negatively charged phospho-
lipid phosphatidylserin (PS), usually exclusively present in the inner lea flet, is transported into the outer lea flet [1
6].
It was shown that the PS-exposure is the key player in the adhesion of RBCs to endothelium cells [72, 73]. It is also
worth to mention that besides e.g. diseases like malaria or s ickle cell disease there are a various number of processes
that are related to PS exposure in the outer lea flet [74, 75, 76 ]. Examples are cell adhesion in general, cell fusion [77],
phagocytosis [78, 70, 79, 80, 81] and apoptosis [82, 79].
3.1. Red Blood Cell Stimulation with LPA
As pointed out in the introductory part of this section, RBCs can be stimulated by LPA, and this has been pro-
posed to contribute to the active participation of RBCs in the later stage of thrombus formation. In order to test for
altered intercellular adhesion behaviour under different conditions, cells were treated with various solution s and were
investigated in micro fluidics with holographic optical twe ezers. Upon stimulation with 2.5 µM LPA, the RBCs ad-
hered to each other within 30 to 60 seconds. During the stimulation procedure, most of the RBCs remained in their
discocyte shape. To exclude any dependencies on the interaction surface due to the anisotropic shape of the cells, we
aimed for another condition using spherocytes [83]. This was realized by increasing the LPA concentration to 10 µM ,
which is still within the physiologically observed range. The separation force could not be determined by the opti-
cal tweezers approach because it exceeds the force of the laser tweezers, which, in this particular setup, amounts to
15 − 25 pN. Consequently, at this stage, the adhesion force could only be qualified to be larger than this. Nevertheless,
with this approach a statement about the general behaviour o f the treated RBCs and the adhesion statistics could be
reached. Cells under five di fferent conditions were tested: (i) in a HEPES buffered solution of physiological ionic
strength (PIS-solution) containing the following (in mM): 145 NaCl,7.5 KCl, 10 glucose and 10 HEPES, pH 7.4, at
room temperature, (ii) PIS-solution containing 2 mM Ca2+ and 10 µM LPA, (iii) PIS-solution containing 2 mM Ca2+
and 2.5µM LPA, (iv) PIS-solution containing 2 mM Ca2+ and no LPA, and finally (v) PIS-solution containing 2 m M
EDTA and 10 µM LPA. We used at least 60 cells per condition. The results are summarized in Fig. 4. The RBC
stimulation with LPA (2.5 µM as well as 10 µM ) in the presence of extracellular Ca2+ led to an immediate qualita-
tive change in the adhesion behaviour: cells stuck irrevers ibly to each other. In case of treatment with 2.5 µM about
72 % of the cells tested showed an irreversible adhesion. In the case of 10 µM , which still represents a physiological
concentration near or inside a blood clot [84], the adhesion rate went even up to more than 90 %. In the control mea-
surement in which the extracellular Ca2+ was chelated by EDTA and in the control measurements without any RBC
treatment, almost no adhesion events could be seen.
8
Figure 4: From ref. [17]. Results of the LPA measurements conducted with optical tweezers. The gray bars represent the pe rcentage of cells that
showed adhesion. The overall number of cells tested was 60 cells per measurement. In the presence of LPA and Ca2+ , a significant number of cells
showed adhesion, whereas in the control experiments, only a very small portion of the cells showed an adhesion. The results of the student’s t-test,
compared to the control measurement (HIS-solution), are indicated at the top of each bar.
3.2. Quantification of the Intracellular Adhesion
To allow a discussion of a physiological (or pathophysiological) relevance of the described adhesion process, one
needs to determine the separation force. As described above , the separation force exceeds the abilities of the HOT.
Therefore, single-cell force spectroscopy [38] was utiliz ed to determine the force. Two different measurements were
conducted: control measurements in which the cells remained untreated, and measurements in which the cells were
treated with a concentration of 2 .5 µM LPA. A concentration of 2.5 µM was chosen for most of the measurements
because it resembles most physiological concentrations. Later on, further measurements with 10 µM LPA were con-
ducted as well to check again for the in fluence of the cell shap e on the measured forces. Fig. 5a shows two example
curves of measurements with 2 .5 µM LPA treated cells (red) and untreated cells (green). In most of the measurements,
a significantly changed adhesion behaviour was observable a fter the treatment with LPA. Whereas the measured forces
in the case of untreated cells barely exceeded 20 pN, in the case of LPA treatment this measured force went up to even
more than 300 pN in some cases. However, the measured forces s trongly varied from cell to cell. In total, more than 50
cells were tested for each case and the results are summarize d in Fig. 5. The mean value of the maximum unbinding
force of untreated RBCs (control, green) amounted to 28 .8 ± 8.9 pN (s.d.) (n=71), whereas in the LPA experiments,
the mean value of the maximum unbinding force amounted to a much higher value of 100 ± 84 pN (s.d.) (n=193, from
three different donors), indicating a severe difference in adhesion behavior of untreated and LPA-stimulated RBCs.
The occurrence of the small adhesion forces in the control me asurements results probably from instrumental artefacts.
Upon stimulation with such LPA, RBCs adhere irreversibly to each other. The separation force of approximately
100 pN (determined by single cell force spectroscopy) is in a range that is of relevance in the vasculature [85]. Due
to the relatively slow response of the RBCs upon stimulation (30 to 60 s) an initiation of a blood clot based on inter-
cellular RBC adhesion is regarded to be irrelevant under phy siological conditions. Once caught in the fibrin network
of a blood clot, the adhesion process observed here in vitro m ay support the solidification of the clot. This notion is
supported by the aforementioned experimental and clinical investigations reporting a prolongation of bleeding time in
subjects with low RBC counts [86, 87, 15, 58].
4. Summary
In this review different adhesion phenomenona of red blood cells (RBC) are discussed. RBC aggregation has
already been known for a long time and is the main determinant of blood viscosity. Additionally, there are many
9
Figure 5: From ref. [17]. a) Shows the combined plot of an example force-distance-curve of a control (green) and an LPA measurement (red). b)
The statistics of measured adhesion forces in the control measurement (without LPA treatment) c) Recorded data of the measured forces for LPA
treated RBCs.
indications that RBC aggregation could play a role in thrombus formation and thrombus solidification. In order to
quantify the different adhesion phenomenona, accurate force measuring tool s are required.
Aggregation that is induced by the presence of plasma macromolecules or in model systems by dextran polymers
is called rouleaux formation. The underlying adhesion mechanism is not fully understood yet. Two different models
were developed to explain the origin of the adhesion, either based on bridging or on depletion. There are experi-
mental data in favour of both theories, but all of them are ind irect measurements and not actual cell-cell interaction
measurements. Cell-cell adhesion measurement of RBCs in their natural, discocytic shape by means of modern spec-
troscopic methods like optical tweezers or atomic force mic roscopy based single cell force spectroscopy (SCFS) were
performed only recently. It turned out that the measured int eraction energies are in excellent agreement with the ones
predicted by the depletion theory. Therefore, based on this data, it can be concluded that the rouleaux formation is
rather depletion-mediated than bridging-mediated. For longer contact times of the cells additional enhanced inter-
action energies could suggest an in fluence of bridging in the later stages of adhesion, but could not be conclusively
con firmed.
Among the intrinsic RBC properties that affect their aggregation, the elastic properties of their memb ranes play
an important role. In order for the RBC aggregates to be stabl e, the cell-cell attraction has to be stronger than the
energy needed for the accompanying RBC shape transformation. Recent corresponding theoretical work emphasized
specificities in treating the adhesion between flexible surf
aces. A significant theoretical prediction was about the rol e
of RBC doublets in the formation of rouleaux.
The platelets induced adhesion of red blood cells has been only recently studied. A combined approach of mi-
cro fluidics and holographic optical tweezers were used to te st the hypothesis statistically. The arising adhesion ex-
ceeded the force capabilities of optical tweezers, thus single cell force spectroscopy was used to quantitatively inves-
tigate this adhesion phenomenon. This approach quantified t he intercellular adhesion to amount to 100 ± 84 pN in
strength. Concerning the question of physiological signifi cance, the results of HOT and SCFS were combined and it
could be concluded that the LPA-induced intercellular adhesion of RBCs is of importance in the later stages of blood
clotting and actively contributes to blood clot solidificat
ion.
10
Appendix A. Acknowledgements
References
[1] R. Hebbel, O. Yamada, C. F. Moldow, H. Jacob, J. G. White, J. W. EatonAbnormal adherence of sickle erythrocytes to cultured vascular
endothelium, J. Clin. Invest. 65 (1980) 154–160.
[2] J. Shelby, J. White, K. Ganesan, P. Rathod, D. Chiu, A microfluidic model for single-cell capillary obstruction by pla smodium falciparum-
infected erythrocytes, PNAS 100 (2003) 14618–14622.
[3] G. Barabino, M. Platt, D. Kaul, Sickle cell biomechanics, Annu. Rev. Biomed. Eng. 12 (2010) 345–367.
[4] O.Baskurt, B. Neu, H. Meiselman, Red Blood Cell Aggregation, CRC Press Taylor and Francis Group, 2012.
[5] S. Chien, L. Jan, Ultrastructural basis of the mechanism of rouleaux formation, Microvasc. Res. 5 (1973) 155–166.
[6] S. Chien, L. Sung, Physicochemical basis and clinical implications of red cell aggregation, Clin. Hemorheol. 7 (1987) 71–91.
[7] R. Fahraeus, The suspension stability of the blood, Physiol. Rev. 9 (1929) 241–274.
[8] M. W. Rampling, Rouleaux formation -its causes, estimation and consequences, D. Tur. J. Med. Sci. 14 (1990) 447–453.
[9] H. Schmid-Sch onbein, R. Wells, R. Schildkraut, Microscopy and viscometry of blood flowing under uniform shear rate , J. Appl. Physiol. 26
(1969) 674–678.
[10] SGerrit Danker, Thierry Biben, Thomas Podgorski, Claude Verdier, and Chaouqi Misbah, Dynamics and rheology of a dilute suspension of
vesicles: Higher-order theory, Phys. Rev. E 76 (2007) 041905–041915.
[11] M. Carr, Y. Hauge, Enhancement of red blood cell washout from blood clots by alteration of gel pore size and red cell fle xibility, Am.J.
Physiol. Heart Circ. Physiol. 259 (1990) H1527.
[12] W. Duke, The relation of blood platelets to hemorrhagic disease, JAMA 55 (1910) 1185–1192.
[13] D. Andrews, P. Low, Role of red blood cells in thombosis, Curr. Opin. Hematol. 6 (1999) 76.
[14] M. D. Horne, A. Cullinane, P. Merryman, E. Hoddeson, The effect of red blood cells on thrombin generation, Br. J. Haemato l. 133 (2006)
403.
[15] L. Kaestner, W. Tabellion, P. Lipp, I. Bernhardt, Prostaglandin E2 activates chanel-mediated calcium entry in human erythrocytes: an indica-
tion for a blood clot formation supporting process, Thromb. Haemostasis 92 (2004) 1269–1272.
[16] D. Nguyen, Phosphatidylserine exposure in red blood cells: A suggestion for the active role of red blood cells in blood clot formation, Ph.D.
thesis, Saarland University (2010).
[17] P. Steffen, A. Jung, D. Nguyen, T .M uller, I. Bernhardt, L. Kaestner, C. Wagner, Stimulation of human red blood cells leads to Ca2+ -mediated
intercellular adhesion, Cell Calcium 50 (1) (2011) 54–61.
[18] K. Brummel, S. Butenas, K. Mann, An integrated study of fi brinogen during blood coagulation, J. Biol. Chem. 274 (1999) 22862–22870.
[19] A. Pribush, D. Zilberman-Kravits, N. Meyerstein, The mechanism of the dextran-induced red blood cell aggregation , Eur. Biophys. J. 36
(2007) 85–94.
[20] Z. Marton, G. Kesmarky, J. Vekasi, A. Cser, R. Russai, B. Horvath, K. Toth, Red blood cell aggregation measurements in whole blood and in
fibrinogen solutions by di fferent methods, Clin. Hemorheol. Microcirc. 24 (2001) 75–83 .
[21] D. Brooks, Mechanism of red cell aggregation, in Blood Cells, Rheology and Aging, D. Platt (ed.) Springer-Verlag, 1988.
[22] D. Brooks, The effect of neutral polymers on the electrokinetic potential of cells and other charged particles: Iv. electrostatic effects in
dextran-mediated cellular interactions, J. Coll. Interf. Sci. 43 (1973) 714–726.
[23] S. Chien, R. J. Dellenback, S. Usami, D. A. Burton, P. F. Gustavson, V. Magazinovic, Blood volume, hemodynamic and metabolic changes
in hemorrhagic shock in normal and splenectomized dogs, Am. J. Physiol. 225 (1973) 866–879.
[24] P. Snabre, P. Mills, Effect of dextran polymer on glycocalyx structure and cell elec trophoretic mobility, Coll. Polym. Sci. 263 (1985) 494–500 .
[25] S. Chien, Biophysical behavior of red cells in suspensions, in: The Red Blood Cell, Academic Press, 1975.
[26] S. Asakura, F. Oosawa, Interactions between particles suspended in solutions of macromolecules, J. Polym. Sci. 33 (1958) 183–192.
[27] P. DeGennes, Scaling concepts in polymer physics, Cornell University Press, 1979.
[28] Y. N. Ohshima, H. Sakagami, K. Okumoto, A. Tokoyoda, T. Igarashi, K. B. Shintaku, S. Toride, H. Sekino, K. Kabuto, I. Nishio, Direct
measurement of infinitisemal depletion force in a colloid-p olymer mixture by laser radiation pressure, Phys. Rev. Lett . 78 (1997) 3963–3966.
[29] D. Brooks, R. Greig, J. Jansen, Erythrocyte Mechanics and Blood Flow, New York:A.R. Liss, 1980, Ch. Mechanisms of Erythrocyte Aggre-
gation, pp. 119–140.
[30] D. Lominadze, W. L. Dean, Involvement of fibrinogen spec ific binding in erythrocyte aggregation, FEBS Lett. 517 (200 2) 41–44.
[31] J. Janzen, D. Brooks, Do plasma proteins adsorb to red cells?, Clin. Hemorheol. 9 (1989) 695–714.
[32] J. Janzen, D. Brooks, A critical reevaluation of the nonspecific adsorption of plasma proteins and dextrans to eryth rocytes and the role of these
in rouleaux formation, in: M. Bender (Ed.), Interfacial Phenomena in Biological Systems, New York: Marcel Dekker, 1991, pp. 193–250.
[33] J. Armstrong, R. Wenby, H. Meiselman, T. Fisher, The hydrodynamic radii of macromolecules and their effect on red blood cell aggregation,
Biophys. J. 87 (2004) 4259–4270.
[34] H. B aumler, E. Donath, A. Krabi, W. Knippel, A. Budde, H. Kiesewetter, Electrophoresis of human red blood cells and platelets: evidence
for depletion of dextran, Biorheology 33 (1996) 333–351.
[35] B. Neu, H. Meiselman, Depletion mediated red blood cell aggregation in polymer solutions, Biophys. J. 83 (2002) 2482–2490.
[36] K. Buxbaum, E. Evans, D. Brooks, Quantitation of surface affinities of red blood cells in dextran solutions and plasma, Biochem. 21 (1982)
3235–3239.
[37] P. Steffen, C. Verdier, C. Wagner, Quantification of depletion induc ed adhesion of red blood cells, Phys. Rev. Lett. 110 (2013) 018102.
[38] J. Friedrichs, J. Helenius, D. J. Muller, Quantifying cellular adhesion to extracellular matrix components by single-cell force spectroscopy,
Nat. Prot. 5 (2010) 1353–1361.
[39] W. Helfrich, Elastic properties of lipid bilayers: Theory and possible experiments, Z. Naturforsch. C 28 (1973) 693–703.
[40] W. Helfrich, Blocked lipid exchange in bilayers and its possible influence on the shape of vesicles, Z. Naturforsch. C 29 (1974) 510–515.
11
[41] E. Evans, Minimum energy analysis of membrane deformation applied to pipet aspiration and surface adhesion of red b lood cells, Biophys.
J. 30 (1980) 265–284.
[42] S. Svetina, M. Brumen, B. Zek s, Lipid bilayer elasticity and the bilayer couple inte rpretation of red cell shape transformations and lysis, Stu d.
Biophys. 110 (1985) 177–184.
[43] L. Miao, U. Seifert, M. Wortis, H. G. D obereiner, Budding transitions of fluid-bilayer vesicles: The e ffect of area-difference elasticity, Phys.
Rev. E. 49 (1994) 5389–5407.
[44] W. C. Hwang, R. E. Waugh, Energy of dissociation of lipid bilayer from the membrane skeleton of red blood cells, Biophys. J. 72 (1997)
2669–2678.
[45] U. Seifert, Configurations of fluid membranes and vesicl
es, Adv. Phys. 46 (1997) 13–137.
[46] S. Svetina, Vesicle budding and the origin of cellular life, Chem. Phys. Chem. 10 (2009) 2769–2776.
[47] H. Lim, M. Wortis, R. Mukhopadhyay, Stomatocyte-discocyte-echinocyte sequence of the human red blood cell: Evidence for the bilayer-
couple hypothesis from membrane mechanics, Proc. Natl. Acad. Sci. U S A 99 (2002) 16766–16769.
[48] U. Seifert, R. Lipowsky, Adhesion of vesicles, Phys. Rev. A 42 (1990) 4768–4771.
Zek s, Equilibrium shapes of erythrocytes in rouleau forma tion, Biophys. J. 84 (2003) 1486–1492.
[49] J. Derganc, B. Bo zi c, S. Svetina, B.
[50] M. Deserno, M. M. Mueller, J. Guven, Contact lines for flu id surface adhesion, Phys. Rev. E 76 (2007) 011605.
[51] A. Agrawal, Mechanics of membrane-membrane adhesion, Math. Mech. Solids 16 (2011) 872–886.
[52] R. Skalak, P. Zarda, K. Jan, S. Chien, Mechanics of rouleau formation, Biophys. J. 35 (1981) 771–781.
[53] P. Ziherl, S. Svetina, Flat and sigmoidally curved contact zones in vesicle-vesicle adhesion, Proc. Natl. Acad. Sci. USA 104 (2007) 761–765.
[54] T. Kirschkamp, H. Schmid-Sch onbein, A. Weinberger, R. Smeets, Effects of fibrinogen and a2-macroglobulin and their apheretic elimination
on general blood rheology and rheological characteristics of red blood cell aggregates, Therap. Apher. Dial. 12 (2008) 360–367.
[55] S. Svetina, P. Ziherl, Morphology of small aggregates of red blood cells, Bioelectrochem. 73 (2008) 84–91.
[56] R. Mukhopadhyay, H. G. Lim, M. Wortis, Echinocyte shapes: Bending, streching, and shear determine spicule shape and spacing, Biophys.
J. 82 (2002) 1756–1772.
[57] D. Kuzman, S. Svetina, R. E. Waugh, B. Zek s, Elastic properties of the red blood cell membrane tha t determine echinocyte deformability,
Eur. Biophys J. 33 (2004) 1–15.
[58] N. Mackman, Triggers, teargets and treatments for thrombosis, Nature 451 (2008) 914–918.
[59] S. Chung, O. Bae, K. Lim, J. Noh, M. Lee, Y. Jung, J. Chung, Lysophosphatidic acid induces thrombogenic activity through phosphatidylser-
ine exposure and procoagulant microvesicle generation in human erythrocytes, Arterioscler. Thromb. Vasc. Biol. 27 (2 007) 414–421.
[60] P. Christophersen, P. Bennekou, Evidence for a voltage-gated, non selective cation channel in the human red cell membrane, Biochim.
Biophys. Acta 1065 (1991) 103–106.
[61] L. Kaestner, C. Bollensdorff, I. Bernhardt, Non-selective voltage-activated cation channel in the human red blood cell membrane, Biochim.
Biophys. Acta 1417 (1999) 9–15.
[62] L. Kaestner, I. Bernhardt, Ion channels in the human red blood cell membrane:
Bioelectrochem. 55 (2002) 71–74.
[63] L. Kaestner, W. Tabellion, E. Weiss, I. Bernhardt, P. Lipp, Calcium imaging of individual erythrocytes: Problems and approaches, Cell
Calcium 39 (2006) 13–19.
[64] G. Gardos, The function of calcium in the potassium permeability of human erythrocytes, Biochem. Biophys. Acta 30 (1958) 653–654.
[65] Q. Li, V. Jungmann, A. Kiyatkin, P. Low, Prostaglandin E2 stimulates a Ca2+ - dependent K + channel in human erythrocytes and alters cell
volume and filterability, J. Biol. Chem. 271 (1996) 18651–18
656.
[66] P. Lang, S. Kaiser, S. Myssina, T. Wieder, F. Lang, S. Huber, Role of Ca2+ - activated K + channels in human erythrocyte apoptosis, Am. J.
Physiol. Cell Physiol. 285 (2003) 1553–1560.
[67] F. Basse, J. G. Stout, P. J. Sims, T. Wiedmer, Isolation of an erythrocyte membrane protein that mediates ca2+ -dependent transbilayer
movement of phospholipid, J. Biol. Chem. 271 (1996) 17205–1 7210.
[68] P. Williamson, A. Kulick, A. Zachowski, R. Schlegel, P. Devaux, Ca2+ induces transbilayer redistribution of all major phosphol ipids in human
erythrocytes, Biochem. 31 (1992) 6355–6360.
[69] P. Williamson, E. Bevers, E. Smeets, P. Comfurius, R. Schlegel, R. Zwaal, Continus analysis of the mechanism of activated transbilayer lipid
movement in platelets, Biochem. 34 (1995) 10448–10455.
[70] L. Woon, J. Holland, E. Kable, B. Roufogalis, Ca2+ sensitivity of phospholipid scrambling in human red cell ghosts, Cell Calcium 25 (1999)
313–320.
[71] D. Dekkers, P. Comfurious, E. Bevers, R. Zwaal, Comparison between Ca2+ -induced scrambling of various fluorescently labelled lipi d
analogues in red blood cells, Biochem. J. 362 (2002) 741–747 .
[72] C. Closse, J. Dachary-Prigent, M. Boisseau, Phosphatidylserine related adhesion of human erythroctyes to vascular endothelium, Br. J.
Haematol. 107 (1999) 300–302.
[73] A. Manodori, G. Barabino, B. Lubin, F. Kuypers, Adherence of phosphatidylserine-exposing erythrocytes to endothelial matrix throm-
bospondin, Blood 95 (2000) 1293–1300.
[74] V. Luvira, S. Chamnanchamnunt, V. Thanachartwet, W. Phumratanaprapin, A. Viriyavejakul, Cerebral venous sinus thrombosis in severe
malaria, Southeast Asian J. Trop. Med. Public Health 40 (2009) 893–897.
[75] A. Eldor, E. Rachmilewitz, The hypercoagulable state in thalassemia, Blood 99 (2002) 36–43.
[76] A. Taher, Z. Otrock, M. Cappellini, Thalassemia and hypercoagulability, Blood Rev. 22 (2008) 283–292.
[77] E. Tullius, P. Williamson, R. Schlegel, Effects of transbilayer phospholipid distribution on erythro cyte fusion, Biosci. Rep. 9 (1989) 623–633.
[78] R. Zwaal, A. Schroit, Pathophysiologic implications of membrane phospholipid asymmetry in blood cells, Blood 89 (1997) 1121–1132.
[79] R. Zwaal, P. Comfurius, E. Bevers, Surface exposure of phosphatidylserine in pathological cells, Cell. and Mol. Life Sci. 62 (2005) 971–988.
[80] Y. Tanaka, A. Schroit, Insertion of fluorescent phospha tidylserine into the plasma membrane of red blood cells- recognition by autologous
macrophages, J. Biol. Chem. 258 (1983) 1335–1343.
[81] A. Schroit, J. Madsen, Y. Tanaka, In vivo recognition and clearance of red blood cells containing phosphatidylserine in their plasma mem-
their further investigation and physiological relevance,
12
branes, J. Biol. Chem. 260 (1985) 5131–5138.
[82] U. Messmer, J. Pfeilschifter, New insights into the mechanism for clearance of apoptotic cells, Bioessays 22 (2000) 878–881.
[83] L. Kaestner, P. Steffen, D. Nguyen, J. Wang, L. Wagner-Britz, A. Jung, C. Wagner, I. Bernhardt, Lysophosphatidic acid induced red blood
cell aggregation in vitro, Bioelectrochem. 87 (2012) 89–95 .
[84] T. Eichholtz, K. Jalink, I. Fahrenfort, W. Moolenaar, The bioactive phospholipid lysophosphatidic acid is released from activated platelets,
Biochem. J. 291 (1993) 677–680.
[85] P. Snabre, M. Bitbol, P. Mills, Cell disaggregation behavior in shear flow, Biophys. J. 51 (1987) 795–807.
[86] A. Hellem, C. Borchgrevink, S. Ames, The role of red cells in haemostasis: the relation between haematocrit, bleeding time and platelet
adhesiveness, Br. J. Haematol. 7 (1961) 42–50.
[87] M. Livio, E. Gotti, D. Marchesi, G. Mecca, G. Remuzzi, G. de Gaetano, Uraemic bleeding: role of anaemia and beneficial
transfusions, Lancet 320 (1982) 1013–1015.
effect of red cell
13
|
1908.08631 | 1 | 1908 | 2019-08-23T01:47:06 | Image based cellular contractile force evaluation with small-world network inspired CNN: SW-UNet | [
"physics.bio-ph",
"cs.CV",
"eess.IV"
] | We propose an image-based cellular contractile force evaluation method using a machine learning technique. We use a special substrate that exhibits wrinkles when cells grab the substrate and contract, and the wrinkles can be used to visualize the force magnitude and direction. In order to extract wrinkles from the microscope images, we develop a new CNN (convolutional neural network) architecture SW-UNet (small-world U-Net), which is a CNN that reflects the concept of the small-world network. The SW-UNet shows better performance in wrinkle segmentation task compared to other methods: the error (Euclidean distance) of SW-UNet is 4.9 times smaller than 2D-FFT (fast Fourier transform) based segmentation approach, and is 2.9 times smaller than U-Net. As a demonstration, we compare the contractile force of U2OS (human osteosarcoma) cells and show that cells with a mutation in the KRAS oncogne show larger force compared to the wild-type cells. Our new machine learning based algorithm provides us an efficient, automated and accurate method to evaluate the cell contractile force. | physics.bio-ph | physics | Image based cellular contractile force evaluation
with small-world network inspired CNN: SW-UNet
Li Honghan, Daiki Matsunaga,∗ Tsubasa S. Matsui, Hiroki Aosaki, and Shinji Deguchi†
Division of Bioengineering, Graduate School of Engineering Science, Osaka University, Japan
1-3 Machikaneyama Toyonaka, Osaka, 5608531 Japan
(Dated: August 26, 2019)
9
1
0
2
g
u
A
3
2
]
h
p
-
o
i
b
.
s
c
i
s
y
h
p
[
1
v
1
3
6
8
0
.
8
0
9
1
:
v
i
X
r
a
We propose an image-based cellular contractile force evaluation method using a machine learning
technique. We use a special substrate that exhibits wrinkles when cells grab the substrate and
contract, and the wrinkles can be used to visualize the force magnitude and direction. In order to
extract wrinkles from the microscope images, we develop a new CNN (convolutional neural network)
architecture SW-UNet (small-world U-Net), which is a CNN that reflects the concept of the small-
world network. The SW-UNet shows better performance in wrinkle segmentation task compared
to other methods: the error (Euclidean distance) of SW-UNet is 4.9 times smaller than 2D-FFT
(fast Fourier transform) based segmentation approach, and is 2.9 times smaller than U-Net. As a
demonstration, we compare the contractile force of U2OS (human osteosarcoma) cells and show that
cells with a mutation in the KRAS oncogne show larger force compared to the wild-type cells. Our
new machine learning based algorithm provides us an efficient, automated and accurate method to
evaluate the cell contractile force.
Cellular contractile force is known to regulate diverse
functions, particularly related to cell adhesion, prolifer-
ation and migration, thus acting as an essential driver
in morphogenesis and pathogenesis [1]. Therefore, mea-
suring cellular contractile force is essential to understand
and control the status of living cells. The most com-
mon methods to measure the contractile force are trac-
tion force microscopy (TFM) [2] and microneedle assay
[3, 4]. In TFM, the displacement field is measured by flu-
orescent microbeads embedded inside the substrate, and
the contractile force is evaluated solving the inverse prob-
lem. In microneedle assay, the contractile force is eval-
uated from the deflections of the microneedles on which
cells are plated.
Another method used to evaluate the contractile force
is the wrinkle based measurements [5 -- 9].
In a spe-
cial substrate that has a stiff top layer by heating
[5, 10] or plasma irradiation [8, 9], cells generate wrinkles
when they grab the substrate and contract as shown in
Fig. 1(a)(b), and the wrinkles can be used to visualize the
force magnitude and direction. The wrinkle length can
be used to estimate the force magnitude since the wrinkle
length has a positive correlation with the force strength
[6, 9].
In previous studies, researchers tried to extract
the wrinkles and measure its length manually [6], or by
2D-FFT (fast Fourier transform) based image processing
[8, 9]. Although the wrinkle based measurement provides
a convenient and efficient way to evaluate the contractile
force, it was difficult to extract the wrinkle from the mi-
croscope images both accurately and automatically. In
this work, we proposed a CNN (convolutional neural net-
work) based method to automate segmentation of wrin-
kles from the microscope images.
In recent years, U-Net [11, 12] is widely used in the
segmentation task for biomedical images including those
of cells [13 -- 15]. In this paper, we propose a new CNN
called SW-UNet (small-world U-Net), which is a modi-
fied U-Net that reflects the concept of the small-world
network [16 -- 18]. The small-world network is a network
that has more connection to its neighbouring nodes while
they have less connection to non-neighbouring nodes, and
this attribute can be quantified by SWI (small-world in-
dex) [18]. The original CNN algorithm was initially in-
spired by the neural structure of the striate-cortex from
macaques and cats [19]. Since the attribute of the small-
world network also exists in the neural structure of ani-
mal cortex [20 -- 22], we hypothesize that integrating this
attribute and building SW-UNet will improve the perfor-
mance of CNN. In this work, we built our CNN based on
the structure of U-Net and optimized the connection to
reflect the concept of the small-world network. Although
there are several recent studies [23, 24] worked on image
classifications or recognitions based on the small-world
inspired CNN, our work is one of the first attempts to
work on the image segmentation for a practical appli-
cation. Our work is also important because we provide
comprehensive knowledge how the network structure af-
fects the segmentation performance.
This paper consists of the following four parts. Firstly,
we prepare training datasets for SW-UNet using image
processing techniques. Secondly, we construct the SW-
UNet architecture by importing the attribute of a small-
world network into U-Net. Thirdly, we compare the ac-
curacy of wrinkle extraction with other methods. Finally,
we apply this novel technology to demonstrate that the
contractile force in U2OS (human osteosarcoma) cells is
elevated upon a mutation in the KRAS oncogene.
EXPERIMENTAL MATERIALS
Cell substrate Based on our previous studies [7, 9],
we prepare the substrate that can generate wrinkles re-
2
FIG. 1. Wrinkle generation by the cell contractile force and fabrication method of the substrate: (a) Microscope
images of wrinkles that are generated by the U2OS cell contractile force. The scale bar in the figure has a length of 20
micrometers. (b) Schematic side view of the cell. The contractile forces are generated by cellular endogenous activity, and the
force gives rise to the wrinkle generation. (c) Schematics of our experiment procedures. (i) The PDMS gel layer is coated on
the polystyrene layer. (ii) The oxygen plasma is applied to the PDMS gel layer to oxide the surface layer. (iii) The U2OS cells
are cultured on the substrate.
versibly upon application of cellular forces following steps
as in Fig. 1(c). Firstly, parts A and B of CY 52-276 (Dow
Corning Toray) are mixed at a weight ratio of 1.25:1 to
form a PDMS (polydimethylsiloxane) gel layer that is
coated on a circular cover glass. Secondly, the cover glass
is placed in a 60°C oven for 20 hours to cure the PDMS
gel. Thirdly, oxygen plasma (SEDE-GE, Meiwafosis) is
applied uniformly along the surface of the PDMS layer to
create an oxide layer that works as the substrate for cell
culture. Finally, the substrate is coated with 10 Îijg/mL
collagen type I solution for 3 hours.
Cells U2OS cells (HTB-96; ATCC) were maintained
in DMEM (043-30085; Wako) supplemented with 10%
FBS (SAFC Bioscience), 100 U/mL penicillin, and 100
Âţg/ mL streptomycin (168-23191; Wako). Cells were
maintained in a humidified 5% CO2 incubator at 37°C.
Plasmids The human KRAS wild-type cDNA (Ad-
dgene plasmid #83166, a gift from Dominic Esposito)
and KRAS G12V cDNA (Addgene plasmid #83169, a
gift from Dominic Esposito) were amplified using KOD-
plus-Neo DNA polymerase kit (KOD-401; Toyobo). The
expression plasmids encoding mClover2-tagged KRAS
wild-type and mRuby2-tagged KRAS G12V were con-
structed by inserting the PCR-amplified cDNAs into
the mClover2-C1 vector (Addgene plasmid #54577, a
gift from Michael Davidson) and the mRuby2-C1 vector
(Addgene plasmid #54768, a gift from Michael David-
son). Before seeding two populations of KRAS express-
ing cells onto the gel substrate, cells were transiently
transfected with either mClover2-KRAS wild-type or
mRuby2-KRAS G12V using ScreenFect A (299-73203;
Wako).
METHODS
Overview
We overview our CNN-based wrinkle detection system
in Fig 2. The full process consists of these three steps:
(a)-(b) preparing the training dataset, (c) training and
(d) wrinkle segmentation. Firstly, we utilize 2D-FFT
method [8] and curvature filter [25] to extract rough wrin-
kle images for the CNN training, as shown in Fig. 2(a).
Note images of cells and wrinkles are captured on an in-
verted phase-contrast microscope (IX73; Olympus) using
a camera (ORCA-R2; Hamamatsu) with a 20× objective
lens. A large number of cells cultured on the same sub-
strate were imaged almost simultaneously using an XY
motorized stage (Sigma Koki). In this step, the wrinkles
are detected purely by the image processing techniques,
and image augmentation is used to increase the number of
training data. Secondly, we train SW-UNet using images
that we prepared in the first step: raw cell image (input)
and wrinkle image (label) shown in Fig. 2(c). Finally,
we utilize this SW-UNet to obtain the wrinkles from test
images as in Fig. 2(d). In the following subsections, we
explain each step in detail.
Training dataset preparation
2D-FFT and bandpass filter The wrinkle patterns are
firstly extracted by combinations of successive three op-
erations: 2D-FFT, bandpass filtering and inverse FFT
(IFFT) techniques [8, 9]. Note this approach has been
already established and utilized in our previous studies
3
FIG. 2. Overview of our approach.
(a) Preparation of training dataset. The wrinkles are extracted by image processing
techniques, 2D-FFT (bandpass filtering) and curvature filter. (b) Image augmentation methods, affine and warping transforma-
tion, are used to increase the number of the training dataset. (c) Training SW-UNet from two images: the original microscope
images and extracted wrinkle images. (d) Utilize SW-UNet to extract wrinkles.
[8, 9], and please refer to these papers for details. Since
the wrinkles have a characteristic wavelength (3-6 pix-
els), the pattern can be extracted applying a bandpass
filter to the image after the 2D-FFT operation as shown
in Fig. 3(a). Restoring the image with IFFT, the wrin-
kles can be extracted as the figure, but the image also
contains cell contours.
Curvature filter Curvature filter is originally designed
to achieve efficient smoothing and denoising operations
[25]. Considering the image intensities as a heightfield,
the surface mean curvature can be obtained at each pixel.
The filter can be used to smooth out only wrinkles be-
cause pixels that have higher curvature decay faster in
this filter. Figure 3(b) shows images before and after the
curvature filter, and it is clearly shown that the wrinkles
smoothed out, and only cell contours remained. Note we
utilized the filter repeatedly 200-1000 times until only
wrinkles disappear.
Computing conjunction (A ∩ B) of two resultant im-
ages, A (right end of Fig. 3(a)) and B (right end of (b)),
the cell contours that appear in image A can be extracted.
Finally substituting the cell contours (A∩ B) from image
A as shown in Fig. 3(c), images with only wrinkles are
obtained.
Image augmentation We prepared 126 original cell
images for the training. Many previous researches that
handle biomedical images [11, 26] used image augmenta-
tion techniques to increase the number of training images.
In this study, we also expand the quantity of our cell im-
ages from 126 to 1404 by the geometric affine transfor-
mations [27, 28] and warping transformations.
CNN architecture
Although the traditional image processing techniques
are effective as shown in the previous section, the method
fails to reproduce the wrinkle pattern in some cases (as
also shown later in Fig. 6(a)). This image processing
approach is not applicable in following three situations:
(i) when the wrinkles are entirely underneath and over-
lapped with the cell, (ii) when the wrinkles have fewer
features of wave-like patterns and (iii) when there are in-
tense noises in the images. In this work, we utilize CNN
to overcome the situation and to extract clear wrinkle
images.
In recent researches, U-Net [11] has been widely used
for segmentation of biological and medical images [29 --
31]. Figure 4(b) shows network topology of U-Net, and
each node corresponds to the tensor format (Nx, Ny, Np);
Nx and Ny represent the image size in pixel units both
x- and y-direction respectively, while Np is the number
of images. Starting from a single input image (Nx, Ny,
Np = 1), which is shown with a blue node in Fig. 4(b), the
input image goes through the network counter clockwise.
Lines between the nodes are the tensor conversions, such
as the pooling and convolution operations. The image
would finally come back to a single output image (Nx, Ny,
1) at the green node, and the network is designed to
extract the desired segmented image at this final tensor.
The U-Net mainly consists of two paths, contracting
path (left side of Fig. 4(b)) and expansive path (right
side). The contracting path is responsible for extracting
the feature from the images, while the expansive path is
designed to reconstruct the desired object from the image
features. The contracting path shrinks the image size
4
FIG. 3. Preparation of training data: wrinkle extraction with image processing techniques. (a) Rough extraction of
wrinkles by a combination of three operations: 2D-FFT, bandpass filtering and IFFT. Since the wrinkles have their characteristic
wavelength (3-6 pixels), they can be extracted (bandpass filtering) and restored (IFFT) with these three steps. (b) Extracting
cell contours from the original images utilizing the curvature filter. Smoothing out the wrinkles, which has a smaller wavelength
(i.e. high curvature), the cell contour is extracted. (c) Constructing clear wrinkle image combining two resultant images A and
B.
using the alternate operations of convolution and pooling
in the order of (pooling, convolution, convolution). As
the result of these procedures, Nx and Ny decrease while
Np increases. On the other hand, the expansive path
increases the image size Nx and Ny while decreasing Np
using alternate operations of (upsampling, convolution,
convolution). The image sizes Nx and Ny reach to a
minimal after the contracting path, and come back to the
original size after the expansive path. There are special
bypass connections in U-Net called "copy and crop" path
[11], which goes horizontally from the contracting to the
expansive path in Fig. 4(b), and the path is responsible
for avoiding the loss of effective information during the
operation of pooling.
Algorithm building SW-UNet
We now introduce the concept of the small-world and
modify the CNN topology. The topology of the small-
world network is characterized and controlled by three
parameters: N, K and P [16, 32]: N is the number
of nodes in the network, K is the average number of
the connection branches between the neighbouring nodes,
and P is the random reconnection probability. The to-
tal number of branches is KN/2, and selected ∼ KN P/2
branches are randomly re-connected to other nodes in the
network. Figure 4(a) shows the schematic of the small-
world network topology under fixed N = 8 and K = 4,
but different P parameter. Each node has connections
only to its closest neighbouring K nodes for P = 0, and
the network topology becomes disordered with the in-
crease of P . We built our SW-UNet architecture through
the following procedures.
Network topology generation In the first step, we
build the DenseNet [33 -- 35] with N = 27, K = 4 as
shown in Fig. 4(c). Each node corresponds to a ten-
sor format (Nx, Ny, Np), and the input image would go
through the network counter-clockwise as U-Net. Follow-
ing the tensor conversions of U-Net, SW-UNet also con-
sists of the contracting path with successive operations
of (pooling, convolution, convolution) and the expansive
path with (upsampling, convolution, convolution). Fig-
ure 4(f) shows the list of tensor formats that we use in
5
FIG. 4. Overview of SW-UNet architecture. (a) Network topology difference based on the random re-connection prob-
ability P . (b)-(e) Network topology of several CNN structures. Each node corresponds to tensor format, while black lines
correspond to the tensor conversions. (f) A table showing the node connection status for (e) SW-UNet. Labels on the hor-
izontal and vertical axis are both tensor formats, and the colors inside the table represent the connection status: red shows
connected nodes, blue shows unconnected nodes and orange shows connected nodes but with the recursively reduced number
of input images. (g) A schematic showing a tensor conversion with three input nodes A − C and a single output node D.
SW-UNet.
In the second step, we reconnect randomly selected ∼
N KP/2 connections for P (cid:54)= 0 as shown in Fig. 4(d)-(e).
The network is DenseNet for P = 0, while the network
is totally random for P = 1 as shown in Fig. 4(e). The
image flow direction is always from the upstream to the
downstream node.
Node connection The format conversions are neces-
sary to connect nodes that have different tensor formats,
and Fig. 4(g) is a schematic of our connection algo-
rithm. The extracted connections are from Fig. 4(d),
and it shows a situation that three input nodes A − C
are connected to a single output node D. We first use
the pooling and up-sampling operations to match the im-
age size of destination node D, N D
y = 32. For
x = N D
example, the pooling operation is utilized to contract
large images as node A (N A
y = 256), while up-
sampling operation is utilized to expand smaller images
as node C (N C
y = 16). Summing up all resul-
tant images from node A-C, the number of total images
is now N D(cid:48)
would
not necessary match the destination node image num-
ber N D(cid:48)
. Therefore, the convolution operation is
p = N A
(cid:54)= N D
but the value N D(cid:48)
p + N B
p + N C
p
x = N C
x = N A
p
p
p
p
p
p
utilized to convert the image number from N D(cid:48)
.
to N D
, N B
Note when one of the input image number (N A
p
, we
) exceeds the destination image number N D
and N C
halve the input image number recursively until they be-
. Figure 4(f) shows the connection
come smaller than N D
status for the network P = 0.4 (Fig. 4(e)): red shows con-
nected nodes, blue shows unconnected nodes, and orange
shows connected nodes but with the recursively reduced
number of input images.
p
p
p
Training parameter
The number of the training dataset is 1404 (126 orig-
inal images), and Adam optimizer [36] with a learning
rate of 0.0001 is utilized for training the CNN network.
We used Nvidia Titan Black (2 GPUs) to accelerate the
training process.
In previous studies, researchers prepared original im-
ages in an order of ∼ 1000 [37 -- 41] as the training dataset
to avoid the overfitting. Since we have 126 original im-
ages for the training dataset, we need to restrict our
training epochs [42]. Therefore, we set the training steps
in one epoch as 300 and set the total epochs as 10.
Wrinkle evaluation
After training CNNs, we evaluate its accuracy with
Ntest = 58 test images by comparing with the ground-
truth data. The ground-truth data are produced by three
different researchers that were asked to trace the wrinkle
lines manually. Although the cross-entropy is the stan-
dard method to compare images [11, 39, 43, 44], we did
not use this method because it was not a proper crite-
rion to compare the performance of different networks.
Interestingly, the accuracy (range: 0.9642-0.9759) and
loss (range: 0.798-0.808) in the training process converge
almost to a same value for all networks, though there is a
significant difference in the extracted wrinkles (as shown
in Fig. 5(a)).
Instead, we utilize perimeter length of the wrinkles (cid:96) as
the comparison criteria. In order to obtain the perimeter,
we extract the wrinkle edge with the Prewitt operator at
a threshold of 0.01 and count up the number of edge pix-
els to obtain (cid:96). We introduce two different distances, Eu-
clidean dEU and cosine distance dCOS, to quantify the dif-
ference between the wrinkle perimeter obtained by CNN
(cid:96)CNN and the ground truth (cid:96)GT. Each distance is defined
as
dEU =
− (cid:96)GT
i
)2,
(1)
i
i=1
((cid:96)CNN
(cid:118)(cid:117)(cid:117)(cid:116)Ntest(cid:88)
Ntest(cid:88)
(cid:118)(cid:117)(cid:117)(cid:116)Ntest(cid:88)
i=1
dCOS = 1 −
((cid:96)CNN
i
i
· (cid:96)GT
(cid:118)(cid:117)(cid:117)(cid:116)Ntest(cid:88)
((cid:96)CNN
)2
i
((cid:96)GT
i
)2
)
.
(2)
i=1
i=1
RESULTS
Effect of P -value in SW-UNet
We first evaluate the segmentation performance using
different network topology, SW-Net (P = 0 to 1) and
U-Net, in Fig. 5(a). Although most of the networks suc-
ceeded in extracting the wrinkles to some extent, P = 0
(DenseNet) and P = 1 (SW-UNet) failed, and they only
showed vague regions of wrinkles. Comparing the wrinkle
perimeter length (cid:96) for different SW-UNets, images (i) and
(ii) shows maximum length at intermediate P = 0.4−0.6,
while image (iii) shows larger (cid:96) for larger P -values. For
images (i) and (ii), the wrinkles are well extracted in
P = 0.4 − 0.6 but become less prominent with P in-
crease. As a result, SW-UNets with large P -value would
underestimate the wrinkle length. In the case of image
6
(iii), the network with P = 0.6 − 0.8 overestimates the
wrinkle length because the network failed to distinguish
the cell contours and wrinkles. Figure 5(b) shows the dis-
tance dEU from the manually tracked ground truth, and
the result shows that the segmentation performance is
best at P = 0.2− 0.4. The distance of U-Net was almost
the same as SW-UNet with P = 0.6.
We now introduce SWI (small-world index) [18] to
characterize the network topology, which is defined as
SW I = 1 − (
L − Ll
Lr − Ll
− C − Cr
Cl − Cr
)
(3)
where L is the average path length and C is the clustering
coefficient defined as
1
L =
N (N − 1)
Dij,
i
j(cid:54)=i
N(cid:88)
N(cid:88)
N(cid:88)
N(cid:88)
N(cid:88)
aij) · (
k
j
(
N(cid:88)
i
C =
1
N
aijaihajh
N(cid:88)
.
aij − 1)
(4)
(5)
j
j
D = 1 is the distance between two nodes, N = 27 is
the number of nodes in the network and a is the con-
nection status between two nodes: aij = 1 when nodes
i and j are connected while aij = 0 if the nodes are not
connected. Subscripts l and r describes that the value is
from the regular or random network respectively: Cl and
Cr are the clustering coefficients for regular and random
networks, while Ll and Lr are the average path lengths
in regular and random networks.
Figure 5(b) shows that SWI reaches maximum at
P = 0.2 and gradually decrease with P increase. Plotting
distance dEU as a function of SWI as shown in Fig. 5(c),
the result infers that the network with larger SWI has
better segmentation performance. Note we evaluated the
distance and SWI with three randomly generated net-
work for each P value. In recent years, there was a re-
port on the macaques and cats cortex topology [45], and
the small-world index was estimated as SW I ≈ 0.4 from
their results. The network topology in the brain might be
optimized in the process of evolution. Although we can-
not draw a definite conclusion here because of the small
number of sample data, there is a possibility that the
network SWI is one criterion to judge the performance
when designing a new CNN.
From next sections, we will fix the value to P = 0.4 for
SW-UNet.
Comparison of different segmentation methods
Figure 6(a) compares extracted wrinkles with different
approaches: 2D-FFT based method (image processing
7
FIG. 5. Wrinkle segmentation using SW-UNet with different P -values. (a) Wrinkle images that are produced by
SW-UNet and U-Net, and (cid:96) is the wrinkle perimeter length. (b) Black plots show the distance dEU while red plots show SWI
(3) of the network, as a function of network reconnection parameter P . (c) The distance dEU as a function of SW I, and the
figure indicates that our network SW-UNet achieves better performance for larger SW I.
based segmentation), U-Net and our SW-UNet. The 2D-
FFT based method has the worst segmentation perfor-
mance, and extracted wrinkles are dotted-line-like pat-
terns rather than continuous lines. This is because the
2D-FFT based method can only detect the patterns that
have periodic wave patterns, and it has a limitation de-
tecting complex-shaped wrinkles as images (ii) or (iii).
The third row of Fig. 6(a) shows the images generated
by U-Net. Although the wrinkles are extracted clearer
compared to the 2D-FFT based approach, U-Net failed to
distinguish the cell contours and wrinkles in some circum-
stances. For example, U-Net treated the cell organelles
as the wrinkles in images (ii) and (iii) and accordingly
overestimating the length of wrinkles. In the case of im-
age (iv), U-Net detected wrinkles at the cell perimeter
even though there are no apparent wrinkles in the micro-
scope image. On the other hand, SW-UNet succeeded
in distinguishing the wrinkles from the cell contour, and
the wrinkle length can be evaluated precisely.
We now introduce the Euclidean distance (1) and co-
sine distance (2) to quantify the segmentation accuracy.
Figure 6(b) shows the accuracy, which is the inverse of
the distance 1/d, obtained by comparing with manually
traced wrinkle lines. Note the accuracy 1/d is normalized
by the score of SW-UNet in the figure. The figure shows
that SW-UNet has far better performance compared to
other two approaches, and the accuracy based on Eu-
clidean distance 1/dEU was 4.9 times accurate compared
to the 2D-FFT based approach, and 2.9 times accurate
compared to U-Net. In the case of the accuracy based
on cosine distance 1/dCOS, it was 36.8 times accurate
compared to 2D-FFT based approach, and 5.5 times ac-
curate compared to U-Net. In summary, our SW-UNet
is the most effective method for this application.
Demonstration: Effect of KRAS mutation
To demonstrate that our SW-UNet is applicable to
evaluate the cellular contractile force, we finally evalu-
ate the force with and without a KRAS mutation and
compare them. Mutations in the KRAS oncogene are
highly correlated with various types of cancer develop-
8
FIG. 6. Wrinkle segmentation accuracy of SW-UNet and its application.
(a) Comparison of extracted wrinkles by
different methods. (b) Accuracy of wrinkle segmentation quantified by the distances, Euclidean and cosine distances, from the
ground truth data. SW-UNet has the smallest error compared to 2D-FFT based segmentation and U-Net. (c) The wrinkles
(green lines) extracted from the microscope images by SW-UNet for U2OS cells with mutant KRAS gene (first row), and
wild-type U2OS cells (second row). (d) Wrinkle lengths of the two cell types. The mutant cell has longer wrinkle compared to
the wild- type, and there is a significant difference (student's t-test) in two groups.
ment [46],
including metastatic colorectal cancer [47],
pancreatic cancer [48] and non-small cell lung cancer [49].
G12V, which is a point mutation with a replacement from
glycine to valine at amino acid 12, is one of the most com-
mon oncogenic KRAS mutations and has been reported
to result in enhanced myosin phosphorylation [50].
Utilizing our new SW-UNet method, we extracted
the wrinkles from the microscope images, as shown in
Fig. 6(c), and the mutant group shows more wrinkles
than the wild-type group. In supplemental meterial, we
also show movies of moving cells with extracted wrinkles
(Movie 1 and 2). Figure 6(d) compares the wrinkle length
(cid:96), and the average length of mutant cells ((cid:96) = 2144) is
larger than that of the wild-type ((cid:96) = 901). Student's
t-test shows that the p-value between these two groups
is 0.0245, and thus indicating that the mutant group and
wild-type group are significantly different. The previous
study [50], which reported enhanced myosin phosphory-
lation upon G12V mutation, indirectly suggests an in-
creased force generation during cancer development. In
accordance with this study, our present result demon-
strates that the mutated cells indeed exhibit greater
forces.
Given that comprehensive analyses are often crucial
in the field of cell biology to evaluate, e.g., how muta-
tions in specific oncogenes or administration of specific
drugs result in changes in cellular physical forces, our
system with SW-UNet of high-throughput capability is
potentially useful to more thoroughly evaluate potential
changes in the cellular contractile force upon different
types of molecular perturbations.
CONCLUSION
In this paper, we proposed an image-based cellular con-
tractile force evaluation method using a machine learning
technique. We developed a new CNN architecture SW-
UNet for the image segmentation task, and the network
reflects the concept of the small-world network. The net-
work topology is controlled by three parameters: number
of nodes N, number of connection branches from a single
node to other K and re-connection probability P . Our
network reaches to the maximum segmentation perfor-
mance at P = 0.2 − 0.4, and the result infers that the
networks with larger SWI might have better performance
in the segmentation. Using our SW-UNet, we can ex-
tract the wrinkles clearer than other methods. The error
(Euclidean distance) of SW-UNet was 4.9 times smaller
than 2D-FFT based wrinkle segmentation approach and
was 2.9 times smaller than U-Net. As a demonstra-
tion, we compared the contractile force of U2OS cells and
showed that cells with mutant KRAS gene exhibit larger
force compared to the wild-type cells. Our new machine
learning based algorithm provides us an efficient, auto-
mated and accurate method to compare the cell contrac-
tile force. We believe that our network SW-UNet and
CNN building strategy would be useful for other appli-
cations.
ACKNOWLEDGEMENT
This work was supported by JSPS KAKENHI Grant
Number 18H03518.
∗ [email protected]
† [email protected]
[1] W. J. Polacheck and C. S. Chen, Nature methods 13, 415
(2016).
journal 80, 1744 (2001).
[3] J. L. Tan, J. Tien, D. M. Pirone, D. S. Gray, K. Bhadri-
and C. S. Chen, Proceedings of the National
raju,
Academy of Sciences 100, 1484 (2003).
[4] Z. Liu, J. L. Tan, D. M. Cohen, M. T. Yang, N. J. Sni-
adecki, S. A. Ruiz, C. M. Nelson, and C. S. Chen, Pro-
ceedings of the National Academy of Sciences 107, 9944
(2010).
[5] K. Burton and D. L. Taylor, Nature 385, 450 (1997).
[6] N. Q. Balaban, U. S. Schwarz, D. Riveline, P. Goichberg,
G. Tzur, I. Sabanay, D. Mahalu, S. Safran, A. Bershad-
sky, L. Addadi, et al., Nature cell biology 3, 466 (2001).
[7] S. Yokoyama, T. S. Matsui, and S. Deguchi, Biochem-
ical and biophysical research communications 482, 975
(2017).
[8] T. Ichikawa, M. Kita, T. S. Matsui, A. I. Nagasato,
T. Araki, S.-H. Chiang, T. Sezaki, Y. Kimura, K. Ueda,
S. Deguchi, et al., J Cell Sci 130, 3517 (2017).
[9] S. P. Fukuda, T. S. Matsui, T. Ichikawa, T. Furukawa,
N. Kioka, S. Fukushima, and S. Deguchi, Development,
growth & differentiation 59, 423 (2017).
[10] A. K. Harris, P. Wild, and D. Stopak, Science 208, 177
(1980).
[11] O. Ronneberger, P. Fischer,
in Inter-
national Conference on Medical image computing and
and T. Brox,
[2] S. Munevar, Y.-l. Wang, and M. Dembo, Biophysical
ing 37, 1418 (2018).
9
computer-assisted intervention (Springer, 2015) pp. 234 --
241.
[12] T. Falk, D. Mai, R. Bensch, Ö. Çiçek, A. Abdulkadir,
Y. Marrakchi, A. Böhm, J. Deubner, Z. Jäckel, K. Sei-
wald, et al., Nature methods 16, 67 (2019).
[13] D. A. Van Valen, T. Kudo, K. M. Lane, D. N. Macklin,
N. T. Quach, M. M. DeFelice, I. Maayan, Y. Tanouchi,
E. A. Ashley, and M. W. Covert, PLoS computational
biology 12, e1005177 (2016).
[14] A. Fabijańska, Artificial intelligence in medicine 88, 1
(2018).
(2008).
[15] H. Niioka, S. Asatani, A. Yoshimura, H. Ohigashi,
S. Tagawa, and J. Miyake, Human cell 31, 87 (2018).
[16] D. J. Watts and S. H. Strogatz, nature 393, 440 (1998).
[17] M. D. Humphries and K. Gurney, PloS one 3, e0002051
[18] Z. P. Neal, Network Science 5, 30 (2017).
[19] D. H. Hubel and T. N. Wiesel, The Journal of physiology
195, 215 (1968).
10, 186 (2009).
[20] E. Bullmore and O. Sporns, Nature reviews neuroscience
[21] M. Rubinov and O. Sporns, Neuroimage 52, 1059 (2010).
[22] E. J. Sanz-Arigita, M. M. Schoonheim, J. S. Damoiseaux,
S. A. Rombouts, E. Maris, F. Barkhof, P. Scheltens, and
C. J. Stam, PloS one 5, e13788 (2010).
[23] S. Xie, A. Kirillov, R. Girshick,
preprint arXiv:1904.01569 (2019).
and K. He, arXiv
[24] M. Javaheripi, B. D. Rouhani, and F. Koushanfar, arXiv
preprint arXiv:1904.04862 (2019).
[25] Y. Gong and I. F. Sbalzarini, IEEE Transactions on Im-
age Processing 26, 1786 (2017).
[26] R. Wu, S. Yan, Y. Shan, Q. Dang, and G. Sun, arXiv
preprint arXiv:1501.02876 (2015).
[27] J. Wang and L. Perez, Convolutional Neural Networks
Vis. Recognit (2017).
[28] Z. Gao, L. Wang, L. Zhou, and J. Zhang, IEEE journal
of biomedical and health informatics 21, 416 (2016).
[29] Y. Han and J. C. Ye, IEEE transactions on medical imag-
[30] B. Kayalibay, G. Jensen, and P. van der Smagt, arXiv
preprint arXiv:1701.03056 (2017).
[31] H. Dong, G. Yang, F. Liu, Y. Mo, and Y. Guo, in annual
conference on medical image understanding and analysis
(Springer, 2017) pp. 506 -- 517.
[32] M. Kochen, The small world (Ablex Pub., 1989).
[33] G. Huang, Z. Liu, L. Van Der Maaten, and K. Q. Wein-
berger, in Proceedings of the IEEE conference on com-
puter vision and pattern recognition (2017) pp. 4700 --
4708.
[34] S. Jégou, M. Drozdzal, D. Vazquez, A. Romero, and
Y. Bengio, in Proceedings of the IEEE Conference on
Computer Vision and Pattern Recognition Workshops
(2017) pp. 11 -- 19.
[35] S. Li, M. Deng, J. Lee, A. Sinha, and G. Barbastathis,
[36] D. P. Kingma and J. Ba, arXiv preprint arXiv:1412.6980
Optica 5, 803 (2018).
(2014).
[37] H. R. Roth, L. Lu, A. Seff, K. M. Cherry, J. Hoffman,
S. Wang, J. Liu, E. Turkbey, and R. M. Summers, in
International conference on medical image computing and
computer-assisted intervention (Springer, 2014) pp. 520 --
527.
[38] A. Depeursinge, A. Vargas, A. Platon, A. Geissbuhler, P.-
A. Poletti, and H. Müller, Computerized medical imag-
ing and graphics 36, 227 (2012).
[39] H.-C. Shin, H. R. Roth, M. Gao, L. Lu, Z. Xu, I. Nogues,
J. Yao, D. Mollura, and R. M. Summers, IEEE transac-
tions on medical imaging 35, 1285 (2016).
[40] R. Zhang, Y. Zheng, T. W. C. Mak, R. Yu, S. H. Wong,
J. Y. Lau, and C. C. Poon, IEEE journal of biomedical
and health informatics 21, 41 (2016).
[41] W. B. Sampaio, E. M. Diniz, A. C. Silva, A. C. De Paiva,
and M. Gattass, Computers in biology and medicine 41,
653 (2011).
[42] J. Loughrey and P. Cunningham, in International Con-
ference on Innovative Techniques and Applications of Ar-
tificial Intelligence (Springer, 2004) pp. 33 -- 43.
[43] J. Shore and R. Johnson, IEEE Transactions on informa-
tion theory 26, 26 (1980).
[44] M. Yi-de, L. Qing, and Q. Zhi-Bai, in Proceedings of
2004 International Symposium on Intelligent Multime-
dia, Video and Speech Processing, 2004. (IEEE, 2004)
pp. 743 -- 746.
10
[45] O. Sporns and J. D. Zwi, Neuroinformatics 2, 145 (2004).
[46] N. Tsuchida, A. K. Murugan, and M. Grieco, Oncotarget
7, 46717 (2016).
[47] R. G. Amado, M. Wolf, M. Peeters, E. Van Cutsem,
S. Siena, D. J. Freeman, T. Juan, R. Sikorski, S. Suggs,
R. Radinsky, S. D. Patterson, and D. D. Chang, Journal
of Clinical Oncology 26, 1626 (2008), pMID: 18316791,
https://doi.org/10.1200/JCO.2007.14.7116.
[48] J. Son, C. A. Lyssiotis, H. Ying, X. Wang, S. Hua, M. Lig-
orio, R. M. Perera, C. R. Ferrone, E. Mullarky, N. Shyh-
Chang, et al., Nature 496, 101 (2013).
[49] G. J. Riely, J. Marks, and W. Pao, Proceedings of the
American Thoracic Society 6, 201 (2009).
[50] C. Hogan, S. Dupré-Crochet, M. Norman, M. Kajita,
C. Zimmermann, A. E. Pelling, E. Piddini, L. A. Baena-
López, J.-P. Vincent, Y. Itoh, et al., Nature cell biology
11, 460 (2009).
|
Subsets and Splits
No saved queries yet
Save your SQL queries to embed, download, and access them later. Queries will appear here once saved.